Flow induced crystallization of polyolefins
PROEFSCHRIFT
ter verkrijging van de graad van doctor aan de
Technische Universiteit Eindhoven, op gezag van de
Rector Magnificus, prof.dr.ir. C.J. van Duijn, voor een
commissie aangewezen door het College voor
Promoties in het openbaar te verdedigen
op woensdag 16 januari 2008 om 14.00 uur
door
Luigi Balzano
geboren te Pompeï, Italië
Dit proefschrift is goedgekeurd door de promotoren:
prof.dr. S. Rastogi
en
prof.dr. P.J. Lemstra
Copromotor:
dr.ir. G.W.M. Peters
A catalogue record is available from the Eindhoven University of Technology Library
ISBN: 978-90-386-1199-0
Copyright © 2008 by L. Balzano
Printed at the Universiteitsdrukkerij, Eindhoven University of Technology, Eindhoven.
Cover design: Luigi Balzano and Bregje Schoffelen (Oranje Vormgevers)
The research described in this dissertation was financially supported by the Dutch Polymer Institute.
(DPI) project # 132.
Table of Contents
Summary ......................................................................................................................... 1
Chapter 1 Introduction ...................................................................................................... 3
1.1
Preamble ..................................................................................................... 3
1.1.1 Process‐properties relation ............................................................ 3
1.1.2 Historical survey on polyethylene and polypropylene ................... 4
1.2
Processing of polyolefins ............................................................................ 5
.
1.3
Aim of the thesis ......................................................................................... 7
1.4
Outline of the thesis .................................................................................... 8
1.4.1 iPP melts containing small amount of DMDBS .............................. 8
.
1.4.2 PE melts with a bimodal molecular weight distribution ................ 9
1.4.3 iPP with unimodal molecular weight distribution .......................... 9
1.5
References ................................................................................................. 10
Chapter 2 Flow induced crystallization in iPP‐DMDBS blends: implications on morphology
of shear and phase separation ....................................................................... 13
2.1
Introduction .............................................................................................. 14
2.2
Experimental method ............................................................................... 16
2.2.1 Materials ....................................................................................... 16
2.2.2 Sample preparation ...................................................................... 17
2.2.3 X‐Ray characterization .................................................................. 17
2.2.4 Rheological characterization ........................................................ 19
2.2.5 DSC ............................................................................................... 19
.
2.3
Results and discussion .............................................................................. 20
2.3.1 Effects of DMDBS on structure and morphology of iPP in the solid
state .............................................................................................. 20
2.3.2 Crystallization under quiescent conditions .................................. 22
2.3.3 Morphology of the system in Region II ........................................ 26
i
2.3.4 Rheology of the system in Region II ............................................. 29
2.3.5 Effect of flow on iPP‐DMDBS blends near the gel transition ....... 31
2.3.6 Morphological implications of flow and DMDBS phase separation
on the crystallization of iPP .......................................................... 34
2.4
Conclusions ............................................................................................... 36
2.5
References ................................................................................................. 38
Chapter 3 Thermo‐reversible DMDBS phase separation in iPP: effects on flow induced
crystallization ................................................................................................. 41
3.1
Introduction .............................................................................................. 41
3.2
Experimental method ............................................................................... 43
3.2.1 Materials ....................................................................................... 43
3.2.2 Sample Preparation ...................................................................... 43
3.2.3 X‐Ray Characterization ................................................................. 44
3.2.4 Rheological Characterization ........................................................ 45
3.2.5 DSC ............................................................................................... 45
.
3.3
Results and discussion .............................................................................. 45
3.3.1 Thermoreversibility in the phase diagram ................................... 45
3.3.2 Linear viscoelasticity of the system in Region I‐PS ....................... 52
3.3.3 Crystallization on cooling after flow in Region I‐PS ..................... 53
.
3.4
Conclusions ............................................................................................... 58
3.5
References ................................................................................................. 59
Chapter 4 Crystallization and dissolution of flow induced precursors ............................ 63
4.1
Introduction .............................................................................................. 63
4.2
Experimental method ............................................................................... 65
4.2.1 Synthesis of a bimodal HDPE ........................................................ 65
4.2.2 X‐ray characterization .................................................................. 66
4.3
Results and discussion .............................................................................. 67
4.3.1 Rheological characterization ........................................................ 67
ii
4.3.2 Thermodynamics of flow induced precursors .............................. 70
4.3.3 Flow induced precursors just above the equilibrium melting
temperature ................................................................................. 72
4.4
Conclusions ............................................................................................... 79
4.5
References ................................................................................................. 79
Chapter 5 Precursors, crystallization and melting in sheared bimodal HDPE melts ......... 83
5.1
Introduction .............................................................................................. 84
5.2
Experimental method ............................................................................... 85
5.2.1 Material preparation .................................................................... 85
5.3
Characterization........................................................................................ 85
5.3.1 Rheology ....................................................................................... 85
5.3.2 Small Angle X‐Ray Scattering (SAXS). ........................................... 86
5.3.3 Wide Angle X‐Ray Scattering (WAXS or WAXD). .......................... 86
5.3.4 Shear experiments ....................................................................... 87
.
5.4
Results and discussion .............................................................................. 87
5.4.1 Flow induced precursors above the equilibrium melting
temperature ................................................................................. 87
5.4.2 Stable and relaxing precursors above the equilibrium melting
temperature ................................................................................. 90
5.4.3 Flow induced shishes below the equilibrium melting
temperature: the influence of temperature ................................ 91
5.4.4 Flow induced shishes below the equilibrium melting
temperature: the influence of flow conditions ............................ 93
5.4.5 Separating shish creation from the kebab crystallization ............ 95
5.4.6 Crystallization onset temperature after short term shear ......... 100
5.4.7 Melting of shish kebabs .............................................................. 101
5.5
Conclusions ............................................................................................. 103
5.6
References ............................................................................................... 104
iii
Chapter 6 Metastable structures during fast short term shear ..................................... 107
6.1
Introduction ............................................................................................ 107
6.2
Materials and methods ........................................................................... 109
6.2.1 Materials ..................................................................................... 109
6.2.2 X‐ray characterization ................................................................ 109
6.2.3 Shear experiments ..................................................................... 111
.
6.3
Results and Discussion ............................................................................ 112
6.3.1 Flow conditions in short term shear .......................................... 112
6.3.2 Flow induced precursors during short term shear ..................... 113
6.3.3 Crystallization after short term shear. ....................................... 114
6.4
Conclusions ............................................................................................. 121
6.5
References ............................................................................................... 121
Chapter 7 Conclusions and recommendations ............................................................... 125
7.1
Conclusions / Technology assessment .................................................... 125
7.2
Recommendations for future research ................................................... 126
Samenvatting ................................................................................................................ 129
Acknowledgements ....................................................................................................... 133
Curriculum Vitae ........................................................................................................... 135
iv
Summary
Flow induced crystallization of polyolefins
Polymers are a widespread class of materials that provide an often advantageous
combination of properties. Easy processability and high versatility combined with low costs
make polymers the materials for an increasing number of high-tech and commodity
applications. Semi-crystalline polyolefins are an important class of polymers, produced in
more than 150 million metric tons per year. They are used to make a wide range of products
ranging from fibers with superior mechanical properties to flexible packaging and molded
parts. The properties of these materials are related to the whole history of the material, from
chemistry/catalysis in the reactor and, in particular, to the processing conditions. Nowadays,
there is a growing interest in added value to these products by achieving outstanding
properties such as high stiffness (up to ~150 GPa) for fibers and clarity for injection molded
parts. This demands more thorough studies on the process-properties relation that is not yet
fully understood.
The main objective of this thesis is to enhance the nucleating efficiency of the
polymer by inducing oriented structures possessing good epitaxial matching. The goal is
achieved by developing the oriented structures in polymer melt either by (a) making use of
fillers that self assemble into nano sized fibrils and orient under flow or (b) by the addition of
identical higher molar mass molecules possessing considerably higher relaxation times
compared to the base (matrix) polymer.
In the first part of the thesis, the crystallization of isotactic polypropylene (iPP) in the
presence of 1,3:2,4-bis(3,4-dimethylbenzylidene)sorbitol (DMDBS) is discussed. DMDBS is
a small organic compound with a high melting temperature (~250 °C) used as a nucleating
agent and so-called clarifier for iPP. The nucleating efficiency of this compound, in the low
concentration regime (less than 1 wt%), is very high and leads to very small iPP crystallites
1
that confer clarity to the material. DMDBS can crystallize within the molten polymer matrix
forming a percolated network of nano-fibrils whose surface hosts a large number of tailored
nucleation sites. Because of the epitaxial relation between iPP and DMDBS, iPP lamellae
grow always radially on DMDBS fibrils, i.e. with the crystalline c-axis parallel to the fibril
axis, the so the so-called shish-kebab morphology (rather similar to the well-known food
product). Therefore, the orientation of DMDBS fibrils templates the orientation of iPP
lamellae. Randomly oriented DMDBS fibrils lead to randomly oriented iPP lamellae and
aligned DMDBS fibrils lead to aligned iPP lamellae. A long lasting alignment of DMDBS
fibrils can be obtained deforming their network even above the melting point of the polymer.
Nearly 0.5 wt% of oriented DMDBS fibrils can template very oriented (fiber-like) polymer
morphologies.
In the second part of this thesis, the flow induced crystallization of high density
polyethylene (HDPE) with a bimodal molecular weight distribution is discussed. This
material is an intimate blend of low and high molecular weight polymer chains (LMW and
HMW) synthesized with a new chemistry route as described in the thesis of N. Kukalyekar
(Ph.D. thesis Eindhoven University of Technology, December 2007). Just above the
equilibrium melting temperature (T
141.2 ) of the polymer, the mutually entangled
HMW chains can be stretched with shear and, due to the restricted number of molecular
conformations, nucleate into needle-like crystals. By choosing appropriate flow conditions, a
suspension of shishes (extended chain crystals) can be formed while the nucleation of kebabs
(folded chain crystals) is suppressed because of a too high temperature. With perfect epitaxy
matching and a good state of dispersion, shishes are the ideal substrate for the nucleation of
HDPE lamellae. On cooling after the application of shear at 142
, HDPE lamellae nucleate
using shishes as a heterogeneous substrate and, therefore, with the crystalline c-axis parallel
to the shish direction. Similarly to the case of iPP-DMDBS blends, it is observed that nearly
0.5 wt% of pre-aligned shishes can template very oriented (fiber-like) morphologies.
2
Chapter 1
Introduction
1.1 Preamble
1.1.1 Process‐properties relation
Polymeric materials exhibit an intricate process-properties relationship that links the
properties of the final products to the whole history of the material. Figure 1.1 describes the
connections from synthesis, via processing, to product properties.
Figure 1.1: Flow chart describing the process-properties relationship in semi-crystalline
polymeric materials.
In the last decades, many scientific studies have been devoted to identify relevant
parameters and their role in this relationship. Interdisciplinary efforts have led to the
production of new materials, with advantageous properties, that have replaced traditional
ones (glass, ceramics, metal, wood, …) in many applications and have enabled developments
in new areas, like micro-electronics and biomedical applications. However, some aspects of
the process-properties relation in polymeric materials are not yet fully understood. Their
clarification could lead, eventually, to materials with properties tailored to the application. A
3
Chapter 1
modern shift in industrial paradigms demands to achieve this goal without developing ‘new’
polymers but, instead, making use of ‘old’ polymers that are based on relatively cheap and
readily available monomers1. For many applications, polyolefins are the ideal candidates.
1.1.2 Historical survey on polyethylene and polypropylene
Polyolefins are commodity materials obtained by polymerizing olefins (alkenes-1).
Nowadays, with more than 150 million metric tons2 per year, polyolefins are the most
widespread class of polymers. Polyolefins are inert materials and, when recycled,
environmentally harmless. The number of products based on these materials, from packaging
to ballistic, from structural to biomedical applications, increases every day.
The simplest polyolefin is polyethylene (PE) that is a sequence of ethylene
monomers. Polyethylene was discovered in the 1930s by Fawcett and Gibson at ICI in strong
collaboration with prof. T. Michels of the Free University of Amsterdam who pioneered the
behavior of gases at elevated pressures. The first industrial PE grades were produced by the
English company ICI in 1939. Initially, it was possible to produce only a highly branched and
with low density PE (LDPE). This highly amorphous material, with high toughness, is still
used in today’s packaging applications. A major breakthrough came in the 1950s, when
Ziegler3 and Natta4, 5 (1963 Nobel Prize in Chemistry laureates) discovered organometallic
catalysts capable of synthesizing high density linear polyethylene (HDPE). Because of a
regular chain structure, HDPE can partially crystallize and it exhibits better mechanical
properties. In the 1960s, polyethylene attracted the attention of physicists. Pennings6 and
Keller7 pioneered the formation of elongated crystals (shish-kebabs) in stirred solution and
stressed melts. At the same time, Ward9 found that upon solid state drawing of melt
crystallized HDPE, re-organization of the molecules increases the E-modulus up to 60 GPa.
These studies unveiled the role of the morphology in the properties of semicrystalline
materials, enabling developments in the area of high performance materials from flexible
molecules. At the end of the 1970s, at DSM Research in the Netherlands, Smith and
Lemstra10, 11 invented a process to spin ultra high molecular weight PE (UHMWPE) from a
semi-dilute solution. After drawing, E-moduli of up to 150 GPa could be achieved. One of
the last breakthroughs was at the end of the 1970s, when Kaminsky12 discovered metallocene
4
Chapter 1
catalysts allowing for narrow molecular weight distributions and enhancing the control over
chain structure of homo- and co-polymers.
The second simplest polyolefin is polypropylene (PP) that is obtained by
polymerizing propylene. PP is basically PE with a methyl side group every other carbon atom
in the back-bone. The relative orientation of the side groups in the space (tacticity) is very
important for the properties of the material. Atactic PP (aPP), with randomly distributed side
groups, can not crystallize and is a rubbery material. In contrast, isotactic PP (iPP), with the
side groups consistently on one side, has the necessary long range order required for
crystallization. iPP is a competitor for HDPE because it has a higher melting point and can be
made transparent with the use of clarifying agents. The synthesis of iPP was enabled by
Ziegler-Natta catalysts13 and was performed, for the first time, by the Italian company
Montecatini in 1957.
1.2 Processing of polyolefins
This thesis deals with topics closely related to melt-processing of polyolefins.
Polyolefins are often processed via the molten state, applying flows and temperature
gradients. Melt-processing has the advantage of not involving solvents and can be used to
create complicated shapes. However, with the design of a manufacturing process for
polyolefins and, more in general, for all polymeric materials, one should also consider
parameters like molecular weight (Mw) and molecular weight distribution (MwD). It is well
established that the viscosity of the melt (η scales with Mw according to a power law14-18:
M
.
. Melt-processing is possible only for relatively low molecular weight materials. In
the other cases, more complicated routes are available but, often, they are limited to simple
profiles, mostly fibres and tapes.
Flow during processing enhances the crystallization rate of the polymer by promoting
the formation of nuclei of the crystalline phase19-33. This alters the final morphology of the
polymer and thus the (mechanical, optical, transport, …) properties of the material34,
35
.
Remarkably, the final morphology of the polymer strongly depends on the structures, called
precursors, present in the early stages of crystallization. These precursors are structures with
undetectable degree of crystallinity but with a certain degree of order. Occasionally, because
5
Chapter 1
of local density fluctuations or further growth due to flow, precursors exceed some critical
dimensions and become spontaneous growing crystalline nuclei. Flow induced precursors
(FIPs) can be generated at relatively high temperatures (i.e. around the thermodynamic
melting point) and, for a strong enough flow, exhibit an anisotropic morphology. These
precursors can be quite large and they initiate the growth of ‘shish kebabs’; i.e. anisotropic
crystallites made of a fibrillar core decorated with a stack of lamellae
6, 8, 36
. In some cases,
shish-kebabs can entirely replace the spherulitic assemblies of lamellae that are characteristic
for crystallization in quiescent conditions. This can be advantageous for some polymer
products but, definitely, not for all of them. For instance, shish-kebabs cause a high modulus
and a high strength in fibres10,
weakness (brittleness)
39,40
37, 38
. In contrast, they can be the source of mechanical
in injection-moulded products. Figure 1.2 shows examples of
spherulites and shish-kebabs obtained by crystallizing polyolefins under quiescent and flow
conditions.
Figure 1.2: a) Scanning electron micrograph of a melt crystallized iPP spherulites (courtesy
P.Schmit); b) Optical micrograph of iPP spherulites growing in the melt
(reproduced with permission from Figure 2, page 32 of reference 41); c)
polyethylene shish-kebab in the melt (reproduced with permission from reference
42); d) Polyethylene shish-kebabs forming zip fastener structures (reproduced
with permission from reference 43); e) Multiple shishes crossing the same kebabs
in polyethylene (reproduced with permission from Figure 3 of reference 44).
6
Chapter 1
For some polymers, for instance iPP, nucleation is relatively slow and processing is
accelerated with nucleating agents (NAs) 45-47. NAs have a marked impact on the morphology
of the polymer and, by reducing the size of the crystallites, can improve the mechanical
properties and reduce the haze. When using NAs, their chemical nature, concentration,
dispersion and aspect ratio need to be considered as extra parameters affecting the final
morphology of the polymer. In addition, during flow, the nucleating particles influence the
local distribution of stresses enhancing the orientation in the surrounding molecules. This
phenomenon can be very important in the flow induced crystallization of polymer melts
containing fillers48.
1.3 Aim of the thesis
The aim of this thesis is to identify basic principles for the onset of oriented
morphologies (shish-kebabs) during flow of melt-processable semicrystalline polyolefins.
For polyolefins, the objective is often to achieve the desired properties with melt
processing at low costs. When the desired properties are the result of an oriented morphology,
the goal can be attained by a) tailoring the melt with small amounts of a ‘smart’ additive or b)
with a clever choice of the molecular weight distribution, both in combination with the right
processing conditions (temperature and flow history). In our experimental work, we consider
three systems:
•
iPP containing small amount of 1,3:2,4-bis(3,4-dimethylbenzylidene)Sorbitol
or DMDBS;
•
PE with a bimodal molecular weight distribution;
•
iPP with unimodal molecular weight distribution;
In all cases, molecular weights allowing melt processability are selected.
7
Chapter 1
1.4 Outline of the thesis
1.4.1 iPP melts containing small amount of DMDBS
DMDBS is an additive which is used as a nucleating agent for iPP 49, 50. The formula
is shown in Figure 1.3. The affinity between this additive and the polymer is very high. Only
tiny amounts of DMDBS, less than 1 wt%, cause dramatic changes in the morphology of iPP.
Crystal assemblies can become smaller than the wavelength of the light (~400 nm) and turn
iPP from an opaque to a clear and transparent material 51.
Figure 1.3: Chemical structure of DMDBS.
The polar molecules of DMDBS can dissolve in the molten iPP only at very high
temperatures. On cooling, DMDBS self-assembles, phase separating from the melt, and
forms a percolated network of fibrils 52 whose surface hosts nucleation sites tailored for iPP.
The state of the art regarding the crystallization of iPP in presence of small amount of
DMDBS is described in the Introduction to Chapter 2 and Chapter 3.
In Chapter 2*, the impact of DMDBS on the crystallization of iPP is discussed. In
particular, we address the role of DMDBS fibrils in templating the iPP morphology after flow
(shear) at high temperatures where the viscosity of the melt is low and the relaxation times
are short.
In Chapter 3†, the thermo-reversibility of DMDBS phase separation is studied and the
investigation on the role of DMDBS fibrils in templating the iPP morphology is extended to
higher temperatures.
*
Partially reproduced from: Balzano, L. et al. ‘Flow induced crystallization in iPP-DMDBS blends:
implications on morphology of shear and phase separation’, Macromolecules 2007 (Accepted)
†
Partially reproduced from: Balzano, L. et al. ‘Thermo-reversible DMDBS phase separation in iPP:
effects on flow induced crystallization ’, Macromolecules 2008 (Submitted)
8
Chapter 1
1.4.2 PE melts with a bimodal molecular weight distribution
It is well established that small amounts of high molecular weight chains promote the
formation of shish-kebabs during flow induced crystallization
22, 26, 28, 30, 33, 53-58
. It has been
proposed that the underlying mechanism relies on the enhanced creation of flow induced
precursors with anisotropic morphology due to stretching of the high molar mass chain
network
59-61
. Stretched chains have a high segmental orientation that allows them to
crystallize faster than coiled chains
19, 62
and form fibrillar crystals
21
. To validate this
hypothesis, we use a specially synthesized blend of a low molar mass linear HDPE
containing 7 wt% of high molar mass linear HDPE. Under shear, this material exhibits a high
tendency to generate shish-kebabs. Shish generated at high temperature can be used as a
heterogeneous substrate for the nucleation of the rest of the molecules; exact lattice matching
and good state of dispersion make them the ideal nucleating substrate.
The state of the art of flow induced crystallization, relevant to the work presented in
this thesis, is summarized in the Introduction paragraphs of Chapter 4 and Chapter 5.
In Chapter 4*, the dynamics of flow induced precursors just above the equilibrium
melting point is discussed. This investigation unveils, for the first time, the possibility to
generate a suspension of extended chain shishes only.
In Chapter 5†, the investigation on the nature of shishes just above the equilibrium
melting temperature is expanded and their potential as nucleators for the bulk of the polymer
is systematically explored.
1.4.3 iPP with unimodal molecular weight distribution
In Chapter 4 and 5, the formation of shishes via crystallization of needle-like flow
induced precursors is observed at relatively high temperature after the application of shear.
Precursors are partially disordered assemblies of molecules and a question on whether they
*
Partially reproduced from: Balzano, L. et al. ‘Crystallization and dissolution of flow induced
precursors’, Physical Review Letters 2007 (Accepted)
†
Partially reproduced from: Balzano, L. et al. ‘Precursors, crystallization and melting in sheared
bimodal HDPE melts’, Macromolecules 2008 (Submitted)
9
Chapter 1
can survive during flow arises. In Chapter 6*, this issue is addressed during fast short term
shear.
1.5 References
1.
2.
3.
4.
5.
6.
7.
8.
9.
10.
11.
12.
13.
14.
15.
16.
17.
18.
19.
20.
21.
22.
23.
24.
25.
Warzelhan, V.; Brandstetter, F. Macromol. Symp. 2003, 201, 291-300.
Younkin, T. R.; Connor, E. F.; Henderson, J. I.; Friedrich, S. K.; Grubbs, R. H.;
Bansleben, B. A. Science 2000, 287, 460-462.
Ziegler, K. Angew. Chem. 1952, 64, 323.
Natta, G. J. Polym. Sci. 1955, 16, 143.
Natta, G.; Corradini, P. Atti Accad. Nazl. Lincei, Mem. 1955, 8, (4), 73.
Pennings, J.; Kiel , A. M. Kolloid ZZ Polym 1965, 205, 160.
Keller, A. Phyl. Mag. 1957, 2, 1171.
Binsbergen, F. L. Nature 1966, 211, 516-517.
Capaccio, G.; Ward, I. M. Nature 1973, 243, 130.
Smith, P.; Lemstra, P. J. Journal of Materials Science 1980, 15, 505-514.
Smith, P.; Lemstra, P. J.; Pijpers, J. P. L.; Kiel , A. M. Colloid & Polymer Sci. 1981,
259, 1070-1080.
Kaminsky, W. Journal of Polymer Science Part A: Polymer Chemistry 2004, 42, (16),
3911 - 3921.
Natta, G. Nobel Lecture December 12, 1963.
Doi, M.; Edwards, S. F., The theory of polymer dynamics. Clarendon Press: Oxford,
1986.
Dealy, J. M.; Larson, R. G., Structure and Rheology of Molten Polymers. Hanser
Gardner Pubns: Cincinnati, 2006.
Macosko, C., Rheology : principles, measurements, and applications. VCH:
Weinheim, 1994.
Rubinstein, M.; Colby, R. H., Polymer Physics. Oxford University Press: 2003.
de Gennes, P. J., Scaling concepts in polymer physics. Cornell University Press:
Ithaca, NY, 1979.
Devaux, N.; Monasse, B.; Haudin, J. M.; Moldenaers, P.; Vermant, J. Rheol. Acta
2004, 43, 210-222.
Heeley, E. L.; Maidens, A. V.; Olmsted, P. D.; Bras, W.; Dolbnya, I. P.; Fairclough,
J. P. A.; Terril, N. J.; Ryan, A. J. Macromolecules 2003, 36, 3656-3665.
Janeschitz-Kriegl, H.; Eder, G. Journal of Macromolecular Science, Part B 2007, 46,
591-601.
Jerschow, P.; Janeschitz-Kriegl, H. International Polymer Processing 1997, 12, (1),
72-77.
Kumaraswamy, G.; Issaian, A. M.; Kornfield, J. A. Macromolecules 1999, 32, 75377547.
McHugh, A. J. Polym. Eng. Sci. 1982, 22, 15-26.
Muthukumar, M. Advances in Chemical Physics 2004, 128.
*
Partially reproduced from: Balzano, L. et al. ‘Metastable structures during fast short term shear’,
Macromolecules 2008 (Submitted)
10
Chapter 1
26.
27.
28.
29.
30.
31.
32.
33.
34.
35.
36.
37.
38.
39.
40.
41.
42.
43.
44.
45.
46.
47.
48.
49.
50.
51.
52.
53.
54.
Nogales, A.; Hsiao, B. S.; Somani, R. H.; Srinivas, S.; Tsou, A. H.; Balta-Calleja, F.
J.; Ezquerra, T. Polymer 2001, 42, 5247-5256.
Pogodina, N. V.; Winter, H. H.; Srinivas, S. Journal of Polymer Science: Part B:
Polymer Physics 1999, 37, 3512-3519.
Seki, M.; Thurman, D. W.; Oberhauser, J. P.; Kornfield, J. A. Macromolecules 2002,
35, 2583-2594.
Somani, R. H.; Yang, L.; Hsiao, B. S. Physica A 2002, 304, 145-157.
Van der Beek, M. H. E.; Peters, G. W. M.; Meijer, H. E. H. Macromolecules 2006,
39, (5), 1805 -1814.
Yamazaki, S.; Watanabe, K.; Okada, K.; Yamada, K.; Tagashira, K.; Toda, A.;
Hikosaka, M. Polymer 2005, 46, 1685-1692.
Ziabicki, A.; Alfonso, G. C. Macromol. Symp. 2002, 185, 211-231.
Zuidema, H.; Peters, G. W. M.; Meijer, H. E. H. Macromolecular Theory and
Simulations 2001, 10, (5), 447 - 460.
Schrauwen, B. A. G.; Breemen, L. C. A. v.; Spoelstra, A. B.; Govaert, L. E.; Peters,
G. W. M.; Meijer, H. E. H. Macromolecules 2004, 37, 8618-8633.
Meer, D. W. v. d.; Pukanszky, B.; Vancso, G. J. J. Macromol. Sci.-Phys. 2002, 41,
1105-1119.
Hill, M. J.; Barham, P. J.; Keller, A. Colloid & Polymer Sci. 1980, 258, 1023-1037.
Baastiansen, C. W. M., Oriented structures based on flexible polymers. PhD thesis,
Eindhoven University of Technology: 1991.
Govaert, L. E., Deformation behavior of oriented polyethylene fibers. PhD thesis,
Eindhoven University of Technology: 1990.
Ward, I. M., Structure and properties of oriented polymers. Chapman and Hall:
London, 1997.
Schrauwen, B. A. G., Deformation and failure of semi-crystalline polymers. PhD
thesis, Eindhoven University of Technology: 2003.
Basset, D. C.; Franck, F. C.; Keller, A. Phyl Trans Roy Soc London A 1994, 348, 2943.
Hobbs, J. K.; Miles, M. J. Macromolecules 2001, 34, 353-355.
Hobbs, J. K.; Humphris, A. D. L.; Miles, M. J. Macromolecules 2001, 34, 5508-5519.
Hsiao, B. S.; Yang, L.; Somani, R. H.; Avila-Orta, C. A.; Zhu, L. Physical Review
Letters 2005, 94, 117802.
Binsbergen, F. L.; de Lange, B. G. M. Polymer 1970, 11, (6), 309-322.
Beck, H. N. Journal of Applied Polymer Science 1967, 11, (5), 673 - 685.
Binsbergen, F. L. Polymer 1970, 11, (5), 253-267.
Hwang, W. R.; Peters, G. W. M.; Hulsen, M. A.; Meijer, H. E. H. Macromolecules
2006, 39, 8389-8398.
Thierry, A.; Fillon, B.; Straupe, C.; Lotz, B.; Wittmann, J. C. Progr. Colloid Polym.
Sci. 1992, 87, (31), 28-31.
Shepard, T. A.; Delsorbo, C. R.; Louth, R. M.; Walborn, J. L.; Norman, D. A.;
Harvey, N. G.; Spontak, R. J. Journal of Polymer Science: Part B: Polymer Physics
1997, 35, 2617-2628.
Kristiansen, M.; Werner, M.; Tervoort, T.; Smith, P.; Blomehofer, M.; Schmidt, H.
W. Macromolecules 2003, (36), 5150-5156.
Thierry, A.; Straupe, C.; J.;, W.; Lotz, B. Macromol Symp 2006, 241, 103-110.
Heeley, E. L.; Morgovan, A.; Bras, W.; Dolbnya, I. P.; Gleeson, A. J.; Ryan, A. J.
Phys. Chem. Comm. 2002, 5, 158-160.
Yang, L.; Somani, R. H.; Sics, I.; Hsiao, B. S.; Kolb, R.; Lohse, D. J. Phys. Condens.
Matter 2006, 18, 2421-2436.
11
Chapter 1
55.
56.
57.
58.
59.
60.
61.
62.
Bashir, Z.; Odell, J. A.; Keller, A. Journal of Materials Science 1984, 19, 3713-3725.
Kimata, S.; Sakurai, T.; Nozue, Y.; Kasahara, T.; Yamaguchi, N.; Karino, T.;
Shibayama, M.; Kornfield, J. A. Science 2007, 316, (5827), 1014 - 1017.
Dukovski, I.; Muthukumar, M. Journal of Chemical Physics 2003, 118, (14), 66486655.
Ryan, A. J.; Fairclough, J. P. A.; Terril, N. J.; Olmsted, P. D.; Poon, W. C. K.
Faraday Discussions 1999, 112, (13-29).
Keller, A.; Kolnaar, H. W. H., Flow induced orientation and structure formation.
VCH: New York, 1997; Vol. 18.
Somani, R. H.; Yang, L.; Zhu, L.; Hsiao, B. S. Polymer 2005, 46, 8587-8623.
Ogino, Y.; Fukushima, H.; Matsuba, G.; Takahashi, N.; Nishida, K.; Kanaya, T.
Polymer 2006, 47, 5669-5677.
Coppola, S.; Balzano, L.; Gioffredi, E.; Maffettone, P. L.; Grizzuti, N. Polymer 2004,
45, 3249-3256.
12
Chapter 2*
Flow induced crystallization in iPP‐DMDBS
blends: implications on morphology of shear
and phase separation
Nucleation is the limiting stage in the kinetics of polymer crystallization. In
many applications of polymer processing, nucleation is enhanced with the addition of
nucleating agents. 1,3:2,4-bis(3,4-dimethylbenzylidene) sorbitol or DMDBS is a
nucleating agent tailored for isotactic polypropylene (iPP). The presence of DMDBS
changes the phase behavior of the polymer. For high enough temperatures the system
iPP-DMDBS forms a homogeneous solution. However, in the range of concentration
spanning from 0 to 1 wt% of DMDBS, the additive can phase separate/crystallize above
the crystallization temperature of the polymer, forming a percolated network of fibrils.
The surface of these fibrils hosts a large number of sites tailored for the nucleation of
iPP. The aim of this Chapter is to investigate the combined effect of flow and DMDBS
phase separation on the morphology of iPP. To this end, we studied the rheology of
phase separated iPP-DMDBS systems and its morphology with time resolved Small
Angle X-ray Scattering (SAXS). The effect of flow is studied combining rheology, SAXS
and a short term shear protocol. We found that, with phase separation, DMDBS forms
fibrils whose radius (~5 nm) does not depend on the DMDBS concentration. The
growth of these fibrils leads to a percolated network with a mesh size depending on
DMDBS concentration. Compared to the polymer, the relaxation time of the network is
quite long. A shear flow, of 60 s-1 for 3 s, is sufficient to deform the network and to
produce a long-lasting alignment of the fibrils. By design, lateral growth of iPP
lamellae occurs orthogonally to the fibril axis. Therefore, with crystallization, the preorientation of DMDBS fibrils is transformed into orientation of the lamellae. This
peculiarity is used here to design thermo-mechanical histories for obtaining highly
oriented iPP morphologies after shearing well above the melting point of the polymer
(i.e. without any undercooling). In contrast, when shear flow is applied prior to
DMDBS crystallization, SAXS showed that iPP crystallization occurs with isotropic
morphologies.
*
Partially reproduced from: Balzano, L. et al. ‘Flow induced crystallization in iPP-DMDBS blends:
implications on morphology of shear and phase separation’, Macromolecules 2007 (Accepted)
13
Chapter 2
2.1 Introduction
Morphology control is an important issue in polymer processing as it influences a
broad range of properties of the final products. For instance, mechanical, optical and transport
properties of polymeric materials depend on the size and shape of the crystallites1, 2. It is well
known that thermal and mechanical histories do play an important role in the creation of
these morphological features3, 4 and that additives can also have a remarkable influence2, 5-8.
Nucleating agents are a family of additives used to speed up processing rates of polymers. In
the case of isotactic polypropylene (iPP) a common nucleating agent is a sorbitol derivative:
1,3:2,4-bis(3,4-dimethylbenzylidene)sorbitol or DMDBS. DMDBS is a chiral molecule that
can self-assemble or crystallize within the molten polymer matrix. Self-assembly takes place
because of inter-molecular hydrogen bonds formation. Hydrogen bonds, in this case, work
essentially in one direction and drive the molecules to pile up (see Figure 2.1). This leads to
the unidirectional growth of fibrillar crystals. Elementary DMDBS fibrils, in iPP, have a
diameter of ~10 nm and a length up to several microns. They can also form bundles with a
diameter up to 100 nm.
Figure 2.1: A stack of two DMDBS molecules.
Crystallization of DMDBS within the iPP matrix corresponds to a liquid-solid phase
separation, in the following, referred to as DMDBS crystallization or DMDBS phase
separation. The DMDBS molecule has a special ‘butterfly’ configuration, see Figure 1.3. The
‘wings’ of the molecule (phenyl rings with two methyl groups attached) enable dissolution in
the polymer and, at the same time, are tailored nucleation sites for iPP, while the ‘body’
14
Chapter 2
comprises two moieties: one dictates the geometry of the molecule and the other bears the
polar groups (hydroxyls) for hydrogen bond formation9. Polarity is one of the main features
of DMDBS. In contrast, iPP is a fully apolar molecule. This difference becomes clear and
leads to a rich phase diagram when iPP and DMDBS are compounded together.
Kristiansen et al.10 proposed a monotectic model for this phase diagram where the
eutectic point lies near 0.1 wt% of the additive. In their model, miscibility of the two
molecules is always possible at high temperatures (Region I). They define four concentration
regimes based on different phase transitions occurring during the cooling of a homogenous
mixture. From the application point of view, the most interesting concentration regime
extends from ~0.1 wt% to ~1 wt% of DMDBS where iPP exhibits a high clarity. The phase
diagram, in this concentration range, is schematically shown in Figure 2.2.
Figure 2.2: Schematic phase diagram for iPP-DMDBS mixtures up to ~1 wt% (quantitative
data shown in Figure 2.9).
Cooling a homogeneous mixture (Region I) leads to crystallization of DMDBS before
crystallization of the polymer (Region II) 10. With crystallization, DMDBS forms a percolated
network of fibrils suspended in the polymer matrix. The nucleation sites for the polymer
reside on the surface of this network. The fibrillar arrangement provides a high surface to
volume (S/V) ratio and, therefore, a large number of nucleation sites per unit of volume.
However, S/V alone cannot explain the nucleation ability of DMDBS. Thierry et al.9, 11 and
Fillon et al.11 demonstrated that DMDBS is a good nucleating agent for iPP because of a
good lattice matching between its crystals and the 31 helix of the polymer. The same authors
15
Chapter 2
also define an efficiency scale for nucleating agents, ranging from 0 to 100 %, based on
characteristic crystallization temperatures. Dibenzylidene sorbitol (DBS), a nucleating agent
very similar to DMDBS, was rated at 41 %. Among several nucleating agents, they found that
4-Biphenyl carboxylic acid (2 wt% in iPP) has the highest nucleation efficiency (66%).
The effect of several sorbitol based nucleating agents on quiescent crystallization
kinetics and the morphology of iPP has been widely explored12-14, as was the rheology of
these systems12, 15, 16. Surprisingly, little attention has been paid to the role of sorbitol based
nucleating agents on the crystallization of iPP during or after imposition of a flow, the most
common scenario in processing. A notable exception is the work of Nogales et al.17, 18. They
studied the flow induced crystallization of iPP-DBS compounds after the phase separation of
the additive under well defined conditions, by means of both scattering and imaging
techniques. For 1 wt% of DBS, they observed, during cooling, after application of modest
shear flows (shear rates ranging from 0.1 to 20 s-1 at 170 ºC), the formation of polymer
morphologies characterized by high degrees of orientation.
However, the role of DMDBS phase separation in flow induced crystallization of iPPDMDBS blends is not yet fully clarified and is the topic of this Chapter. The work includes
also the changes in the rheology of the melt, associated to the formation of the DMDBS
fibrillar network, and the flow behavior of this network. The results are based on a
combination of Small Angle X-ray Scattering (SAXS), Dynamic Scanning Calorimetry
(DSC) and Rheology. Four different iPP-DMDBS blends, containing 0, 0.3, 0.7 and 1.0 wt%
of the additive are investigated in quiescent and flow conditions. We address three aspects of
these blends: 1. crystallization without application of flow (quiescent conditions); 2.
influence of flow prior to the crystallization of DMDBS; 3. influence of flow after
crystallization of DMDBS.
2.2 Experimental method
2.2.1 Materials
The iPP used in this work is a commercial homopolymer grade from Borealis GmbH
(Austria), labeled HD120MO, with molecular weight, Mw, of 365 kg/mol and a
16
Chapter 2
polydispersity, Mw/Mn, of 5.4. DMDBS (Millad 3988) was obtained in powder form from
Milliken Chemicals (Gent, Belgium) and used as received.
2.2.2 Sample preparation
The polymer, available in pellets, was first cryo-ground and then compounded with
DMDBS in a co-rotating twin screw mini-extruder (DSM, Geleen) for 10 min at temperatures
ranging from 230 to 250 ºC, the higher the DMDBS concentration the higher the
compounding temperature used. To prevent degradation of both, polymer and additive, this
operation was performed in a nitrogen rich atmosphere. The material obtained was
compression molded with a hot press into films of different thicknesses: 1mm for rheology
and 200μm for X-ray experiments. The compression molding temperature was 220 ºC and the
molding time was 3min. The resulting films were quenched to room temperature and cut in
disk-like samples. Following the same procedure, three blends of iPP with 0.3, 0.7 and 1 wt%
of DMDBS were prepared. For convenience, these three blends are respectively referred to as
B03, B07 and B1 in the text.
2.2.3 X‐Ray characterization
X-ray characterization was performed at the European Synchrotron Radiation Facility
(ESRF) in Grenoble (France). Time resolved Small Angle X-ray Scattering (SAXS)
experiments were performed at beamline BM26/DUBBLE. Scattering patterns were recorded
on a two dimensional gas filled detector (512x512 pixels) placed at approximately 7.1 m
from the sample. Scattering and absorption from air were minimized by a vacuum chamber
placed between sample and detector. The wavelength adopted was λ=1.03 Å. SAXS images
were acquired with an exposure of 5 s and were corrected for the intensity of the primary
beam, absorption and sample thickness. The scattered intensity was integrated and plotted
against the scattering vector, q = (4π / λ)sin(ϑ / 2) where ϑ is half of the scattering angle.
The long period was calculated as Lp = 2π / (qI
MAX
) , where qI
MAX
is the q value corresponding
to the maximum in the scattered intensity. Finally, we defined an integrated intensity as:
q max
II =
∫
I (q )dq where q min and q max are the minimum and the maximum experimentally
q min
17
Chapter 2
accessible q values respectively. Two dimensional SAXS images were also used for the
characterization of anisotropic morphologies. For this purpose, it was necessary to define
three azimuthal regions19. The definitions adopted in the present work are given in Figure 2.3.
Figure 2.3: Anisotropic two dimensional SAXS image with definitions of the azimuthal
intensity regions. Arrow indicates the applied flow direction.
Shear flow experiments in combination with SAXS were carried out in a Linkam
Shear Cell (CSS-450) modified with Kapton windows using a ‘short term shearing’ protocol.
First, samples were annealed at 230 ºC for 3 min to erase the memory of any previous
thermo-mechanical treatment. Next, the temperature was decreased by 10 ºC/min to the
desired test temperature where flow was applied under isothermal conditions. For the purpose
of this Chapter, we limit ourselves to the application of only one shear condition: nominal
shear rate of 60 s-1 for 3 s. Finally, depending on the experimental requirements, the
temperature was either decreased to the room temperature or kept constant.
Wide Angle X-ray scattering (WAXD) experiments were performed separately on
beamline ID11 of the ESRF. The results were used to determine crystallinity and the phases
present in the samples. Two dimensional images were recorded on a Frelon detector. Before
analysis, the scattering of air and of the empty sample holder was subtracted. After radial
integration, the intensity was plotted as a function of the scattering angle 2ϑ . Deconvolution
of the amorphous and crystalline scattered intensities was performed using a sixth order
polynomial to capture the ‘amorphous halo20, 21. The crystallinity index, a measure of the
crystal volume fraction, was calculated as:
XWAXD =
AC
AC + AA
18
⋅ 100
(2.1)
Chapter 2
where, AA and AC are the scattered intensities from the amorphous and the crystalline phases,
respectively.
2.2.4 Rheological characterization
Rheological measurements were performed in the linear viscoelastic regime using a
strain-controlled ARES rheometer equipped with a 2KFRT force rebalance transducer. In all
cases a plate-plate geometry with a diameter of 8 mm was used. Appropriate values of strain
were determined with amplitude sweep tests carried out at 5 rad/s over a broad range of
strains (ranging from 0.01 to 100 %)22. During the study of phase transitions, large strains can
enhance the process and/or affect the morphology23. These effects are minimized by using
strains as low as 0.5 % in the experiments.
2.2.5 DSC
The crystallization behavior of the three binary blends iPP-DMDBS was studied in
quiescent conditions using Dynamic Scanning Calorimetry (DSC). Samples of approximately
2 mg were placed into aluminum pans and tested in nitrogen atmosphere in a Q1000
calorimeter (TA Instruments). The first step in the thermal treatment was always annealing at
230 ºC for 3 min to erase earlier thermo-mechanical histories. Next, samples were cooled to
room temperature at a constant cooling rate of 10 ºC/min.
Before identifying peak positions and determining crystallinity, a linear baseline was
subtracted from the measured heat flow as a function of the temperature. Finally, crystallinity
Te
could be estimated as: X
DSC
0
c
= ΔHc ΔH , where ΔH c =
∫ (dH
dT )dT and ΔH c0 are,
Ts
respectively, the enthalpy of crystallization of the sample and the enthalpy of crystallization
of an ideal 100 % crystalline iPP sample (207.1 J g-1) 24.
19
Chapter 2
2.3 Results and discussion
2.3.1 Effects of DMDBS on structure and morphology of iPP in the solid state
In semi-crystalline polymers, structure and morphology depend on the crystallization
conditions (thermal and mechanical histories). In order to isolate the effects due to the
presence of DMDBS in the solid state, samples were prepared under the same crystallization
conditions i.e. quiescent crystallization with 10 ºC/min. Figure 2.4 reports WAXD integrated
intensities at room temperature for the neat iPP and the blends with DMDBS.
Figure 2.4: WAXD profiles of iPP at room temperature as function of DMDBS concentration.
All samples were prepared in the same conditions, i.e. crystallization from the
melt at 10 ºC/min. Presence of DMDBS induces the broad 117 peak, indicated by
the arrow, that is associated to the formation of γ phase crystals. The crystallinity
index is ~60 % in all cases while the amount of γ phase decreases with DMDBS
concentration. Note that, curves are shifted in the vertical direction for clarity.
The neat iPP shows the typical diffraction peaks of the α crystalline modification.
When the additive is present, although the α form remains prevalent, the crystal structure of
the polymer shows some specific changes. The 111 peak becomes better resolved and a broad
117 reflection appears. This indicates the simultaneous formation of less defected α and small
γ crystals. However, we do not observe significant variation in the WAXD crystallinity
index; in all cases, it lies around 60 %. According to Foresta et al.24, the formation of γ phase
crystals in presence of the nucleating agent can be explained from a thermodynamic point of
view. The nucleating agent shifts the crystallization of the polymer at higher temperatures
20
Chapter 2
where nucleation of γ phase is favored and can compete with nucleation of α phase. The ratio
between γ and α phase crystals, X γ , can be estimated with:
Xγ =
A117
A130 + A117
(2.2)
where A130 and A117 are the areas of the non overlapping parts of the 117 and 130 peaks.
These two peaks were selected because they are the diagnostic reflections of the γ and the α
phase respectively. In the investigated range of concentration, the γ phase content, X γ , is
maximum for B03 ( X γ =0.15) and drops for B07 ( X γ =0.09) and B1 ( X γ =0.08). This drop
is probably related to a faster α nucleation rate at higher DMDBS concentrations.
On the morphological side, the long period of iPP lamellae shows pronounced
changes as a function of DMDBS concentration going from 19 nm of the neat sample to 23
nm (average value) of samples containing DMDBS, see Figure 2.5.
Figure 2.5: Long periods of iPP lamellae at room temperature as a function of DMDBS
concentration. All samples were prepared under the same conditions, i.e.
crystallization from melt at 10 ºC/min. The neat iPP shows a long period of 19
nm and this value rises to ~23 nm for samples containing DMDBS. This increase
in long period is due to the formation of thicker crystals in presence of DMDBS
The lamellar thickness, TL, can be expressed as TL = Lp ⋅ x . Since, the crystallinity
index does not vary, our experimental observations are consistent with the formation of
thicker crystals when DMDBS is present. The reason for this increase in crystal thickness is
21
Chapter 2
the higher crystallization temperature in presence of the nucleating agent8 that is discussed
hereafter.
2.3.2 Crystallization under quiescent conditions
When cooling a homogeneous mixture of iPP and DMDBS to room temperature, two
phase transitions are observed: crystallization of DMDBS and crystallization of the polymer.
DSC experiments reveal the temperatures and enthalpies characterizing both these transitions.
In the cooling thermograms of Figure 2.6 the crystallization peaks of the polymer are, in all
cases, clearly visible. A closer look discloses another, much smaller, exotherm at higher
temperatures.
Figure 2.6: DSC cooling thermograms (after subtraction of a linear baseline) for the neat
polymer and blends B03, B07 and B1. Experiments were performed at 10ºC/min,
in N2 atmosphere, after annealing the samples at 250 ºC for 3 min. Curves are
shifted along the vertical axis for clarity. With the addition of only 0.3 wt% of
DMDBS, the crystallization peak shifts to 132 ºC. and its position does not
change with further addition of the additive. Nevertheless, the crystallization peak
becomes narrower when increasing the amount of DMDBS.
This smaller exotherm is associated with the crystallization of DMDBS and, due to the small
amount of the additive, becomes visible only after sufficient magnification, see Figure 2.7.
Some relevant DSC data during the cooling experiments are summarized in Table 2.1. Note
that these data provide enough information to sketch the phase diagram of the system in the
investigated range of concentration.
22
Chapter 2
Figure 2.7: Magnification of the cooling experiments of Figure 2.6 in the temperature range
preceding the crystallization of the polymer. The small exotherms are associated
to the crystallization of DMDBS. As expected, latent heat of crystallization and
peak temperature increase with DMDBS concentration. For clarity, curves are
shifted to the same baseline.
Table 2.1: Summary of experimental data obtained from DSC data shown in Figure 2.6.
DSC
DSC
Tpeak and Tonset represent peak and onset temperature of the exotherm associated
to crystallization of the polymer. tc is the crystallization time defined as
DSC
DSC
DSC
tc = (Tonset −Tcompl ) / dT dt where Tcompl corresponds to the completion of the
DSC
crystallization and dT dt is the cooling rate (=10 ºC/min). Tps represents the
peak temperature of the exotherm associated to DMDBS crystallization. X DSC is
the degree of crystallinity of the polymer.
DSC
Tpeak
DSC
Tonset
[ºC]
HD120MO
DSC
Tps
X DSC
[ºC]
∆H
[J·g-1]
tc [s]
113
120
95.3
123.5
0.3% DMDBS – B03
131
135
107.7
68.5
149
52
0.7% DMDBS – B07
132
135
107.7
53.6
175
52
1% DMDBS – B1
132
135
103.5
47.3
189
50
[ºC]
[%]
46
Upon addition of 0.3 wt% of DMDBS, the crystallization temperature (peak value) of
iPP, Tc, increases to 131 ºC. Further addition of DMDBS has nearly no effect on Tc that is
132 ºC for both B07 and B1. Nevertheless, the crystallization peak of the polymer narrows at
higher DMDBS contents indicating faster crystallization. Saturation of Tc of iPP with
DMDBS concentration was observed also by Kristiansen et al.10, in their data, Tc reaches
~130 ºC at 0.4 wt% DMDBS. Increasing DMDBS concentration, the phase separation occurs
23
Chapter 2
at increasingly higher temperatures. In accordance with WAXD, the final crystallinity of iPP
is hardly affected by DMDBS. However, the values measured by DSC, namely 50 %, are
noticeably lower than those found with WAXD.
Information on the morphology of the system as a function of the temperature is
obtained by means of SAXS. Figure 2.8 shows the integrated scattered intensity as a function
of the temperature for the neat iPP and the blends with DMDBS. These data can be
interpreted in terms of density fluctuations. As expected, in the neat iPP there is no density
fluctuation until the polymer starts nucleating at ~120 ºC. While, samples containing
DMDBS show more complicated temperature dependence. In fact, when phase separation
occurs, DMDBS molecules form crystals denser than the polymer.
Figure 2.8: Temperature dependence of the SAXS intensity as a function of DMDBS
concentration during cooling at 10 ºC/min and after annealing at 250 ºC for 3
min. In samples containing DMDBS the scattered intensity increases with phase
separation because of density fluctuations between DMDBS crystals and the
polymer. At lower temperatures, when the polymer crystallizes once again the
scattered intensity increases
As a consequence, electron density fluctuations are established and the scattered
intensity rises to a plateau. At lower temperature, around 135 ºC, independently from
DMDBS concentration, nucleation of the polymer triggers a large and abrupt upturn in the
intensity. Similar to DSC, some characteristic temperatures for the crystallization of the
polymer and of the additive are located and reported in Table 2.2. These data are used to
build the phase diagram shown in Figure 2.9 that is used as reference in the rest of this work.
24
Chapter 2
In accordance with Kristiansen et al.10, three different regions, corresponding to three
different physical states of the system, are identified:
Region I: at high temperatures DMDBS and iPP form a homogenous solution;
Region II: at intermediate temperatures the system is phase separated with DMDBS
crystallized and iPP molten;
Region III: at low temperatures both DMDBS and iPP are crystallized.
SAXS
Table 2.2: Summary of the SAXS data obtained from Figure 2.9. TcSAXS and Tpeak are,
respectively, the onset temperature for polymer crystallization and the
SAXS
temperature corresponding to the maximum scattered intensity. Tonset ps is the onset
SAXS
temperature for DMDBS phase separation, and Tplateau is the temperature at
which the intensity reaches a constant value (above Tc).
SAXS
Tplateau
TcSAXS
SAXS
Tpeak
SAXS
Tonset ps
[ºC]
[ºC]
[ºC]
HD120MO
120
[ºC]
108
0.3% DMDBS – B03
135
125
165
150
0.7% DMDBS – B03
135
125
190
175
1% DMDBS – B1
135
127
195
185
Figure 2.9: Phase diagram of the system iPP-DMDBS (from 0 to 1 wt% DMDBS) obtained,
on cooling, using SAXS data. Three regions corresponding to three different
states can be identified: Region I) homogeneous liquid, Region II) phase
separated system with crystallized DMDBS and molten polymer, Region III) both
iPP and DMDBS are crystallized
25
Chapter 2
When the polymer crystallizes, in Region III, SAXS allows for the measure of the
long period. Figure 2.10 shows the data concerning the neat polymer, B03, B07 and B1 as a
function of temperature. As already discussed, the presence of DMDBS leads to an increase
in Lp.
Figure 2.10: Long period as a function of temperature and DMDBS concentration during
temperature ramps with cooling rate of 10 ºC/min. Presence of DMDBS leads to
an increase of the long periods that below 80 is quantified in ~4 nm
2.3.3 Morphology of the system in Region II
Two dimensional SAXS images reveal that the increase of the integrated intensity in
Region II is caused by an increase of the scattering in all azimuthal directions at low q.
Sample images are shown in Figure 2.11.
Figure 2.11: SAXS images of the blend B1. Left: material in Region I of the phase diagram
and Right: material in Region II of the phase diagram. DMDBS phase separation
causes an increase of the scattered intensity in all directions at low q values. For
a clear visualization, the scattering of the system in Region I was subtracted.
26
Chapter 2
Such a scattering pattern can be ascribed to the formation of a suspension of randomly
oriented DMDBS fibrillar crystals with a length L and a radius R. In this case, the intensity
scattered in the q region 2π / L < q < 1 / Rc , can be described with25-27:
⎛ R 2q 2 ⎞
C
⎟
⎜
⎟
⎜
I (q ) = Exp ⎜− c ⎟
⎟
⎜
⎟
q
2 ⎠
⎝
(2.3)
where C is a constant including details on the scatterers like concentration and electron
density, while Rc is the radius of gyration of the cross section of the scatterers ( Rc = R / 2
). From the existing literature, it is known that DMDBS fibrils are basically endless (L ∞),
therefore Equation (2.3) is valid for
1/
in this case. Within this limit, Log[I (q ) ⋅ q ]
versus q 2 is a straight line with a slope −Rc2 / 2 . Fitting Equation (2.3) to the data points
allows for the calculation of Rc and therefore of R. Figure 2.12 provides an example of such a
fit demonstrating that a good agreement between experimental data and Equation (2.3) exists
for 0.15100s
-1
15s / 4s
0.68
2500
0.86
67.8
-1
30s / 2s
1.37
5000
2.1
42.8
60s-1 / 1s
2.73
10000
3.7
41
Similar to what is observed at 142
[%]
λ [s]
, with the formation of needle-like FIPs,
rises quickly to a maximum and then decreases (see Figure 5.7). However, in this case the
drop of I SAXS is relatively small and does not correlate with τHMW . Due to the lower
D
temperature, the early FIPs are more stable than at 142
and their lifetimes t
IP
τHMW .
D
Moreover, the drop of I SAXS could also be (partially) explained with the presence of the
kebabs that decrease the density fluctuation between the shishes and the melt.
94
Chapter 5
Figure 5.7: Integrated meridional and equatorial SAXS intensities at 137
for 4 s.
15
after a shear of
5.4.5 Separating shish creation from the kebab crystallization
We are now going to show how shishes nucleate the bulk of the polymer. In Chapter
4, the formation of crystalline shishes with shear flow at 142
is discussed. Just like
nucleating agents, shishes generated at this high temperature can be used as a heterogeneous
substrate for the nucleation of the rest of the molecules at a lower temperature. An exact
lattice matching and a good state of dispersion make shishes the perfect nucleating agent. In
this paragraph, we discuss the increase in the crystallization temperature and the changes in
. Different flow conditions
the morphology induced by shishes generated with shear at 142
are investigated and the cooling starts immediately after flow. Figure 5.8 shows SAXS
images taken at 142
shortly after the application of shear. The corresponding De number
for HMW molecules are given in Table 5.2. The scattering fingerprint of needle-like FIPs and
shishes is clear, independently of shear rate, when γ=100. In all the other cases, in the
accessible q range, no changes are observed in the scattering patterns immediately after flow.
Table 5.2: Deborah numbers for HMW chains in shear at 142
-1
-1
-1
-1
.
2s
5s
10 s
25 s
50 s-1
100 s-1
HMW Des
0.07
0.2
0.4
0.9
1.8
3.5
HMW Deo
260
650
1300
3250
6500
13000
95
Chapter 5
The scattering patterns of Figure 5.9 suggest the presence of a threshold strain γ
for shish formation. Below γ
, regardless of , the stretch ratio of the HMW molecules
would be insufficient to form shishes. This idea is consistent with the findings of van
Meerveld et al.37.
However, the absence of an equatorial streak in SAXS does not,
necessarily, imply the absence of FIPs and shishes. The equatorial streak is absent also when
the concentration of FIPs and shishes is (too) low. This is proven, during cooling, by the
growth of lamellae (kebabs) with a high degree of orientation (see Figure 5.9), suggesting the
presence of an anisotropic nucleating substrate (FIPs or shishes). According to the
simulations of Frenkel and coworkers19, even a single stretched molecule can nucleate
parallel lamellae. Surprisingly, in the range of conditions explored in this Chapter, the overall
degree of lamellar orientation is found to depend mainly on γ. Very similar orientations are
observed at a given γ, independently of . While, much more pronounced differences are
observed varying γ (or varying t ) at a given . Figure 5.9 demonstrates that the influence of
shear flow on PE morphology is substantial and can be dramatic in some cases. For instance,
at γ=100, starting from
25
, isotropic scattering is hardly detectable. For these
conditions, all lamellae grow with the c-axis parallel to the shishes (i.e. parallel to the flow
direction) yielding a very high level of anisotropy. The emergence of a second order
reflection at the meridian of SAXS (see patterns in Figure 5.9) indicates that kebabs are
distributed along the flow direction with high regularity. At room temperature, their spacing
is 23 nm. Morphology evolution during bulk crystallization is discussed for the case of
cooling after a pulse of 100 s-1 for 1 s at 142
, see Figure 5.10. At high temperature, the
SAXS equatorial intensity is higher than the meridian and diagonal due to the scattering of
FIPs and shishes. During cooling, L
dissolve at 142
and D
become smaller and, small FIPs, bound to
, can crystallize (thermal nucleation) increasing the equatorial intensity.
Simultaneously, also the meridian intensity increases but, no maximum can be observed in
the intensity profile above T 130 , suggesting uncorrelated objects in the melt. In other
words, because of thermal nucleation, small (point-like) FIPs are transformed in stable nuclei.
These small FIPs can tumble during shear and, therefore, have random orientations.
Moreover, the orientation of these small FIPs is lost already at a relatively small lateral
growth.
96
Chapter 5
Figure 5.8: SAXS images few seconds after the application of shear at 142
direction is horizontal. The shear time is indicated in each image.
97
. The flow
Chapter 5
Figure 5.9: SAXS images at room temperature after the application of shear at 142
flow direction is horizontal. The shear time is indicated in each image.
98
. The
Chapter 5
Figure 5.10: SAXS equatorial and meridional integrated intensities as a function of
temperature after shear at 142 (cooling rate is 5 /min).
At T 133
, large scale (secondary) nucleation starts. First, the heterogeneous
nucleation of kebabs on the surface of shishes causes an increase in I SAXS and, next, at lower
temperatures, some isotropic nucleation in the bulk of the sample causes an increase in I SAXS
as well. Particularly interesting is the case of mixed morphologies where spherulites coexist
with shish-kebabs. In these cases, it is possible to distinguish between the high temperature
nucleation of kebabs from the low temperature nucleation of spherulites. An example is the
experiment shown in Figure 5.11 (cooling after 10 s
for 5 s at 142
).
Figure 5.11: Salient SAXS images during crystallization on cooling after shear at 142
99
.
Chapter 5
5.4.6 Crystallization onset temperature after short term shear
As for nucleating agents, the nucleation ability of shishes can be estimated comparing
the temperatures where heterogeneous nucleation occurs during cooling. We found that the
onset temperature of the kebabs T
depends on the flow conditions at 142
conditions described in the previous paragraph, T
varies from ~125
. For the flow
to ~132
, as
reported in Figure 5.12.
Figure 5.12: Onset temperature for the nucleation of kebabs on cooling (5
shear at 142 .
Similar to the morphological orientation, T
and on t . Consistent with the SAXS data, T
/min) after
depends mostly on γ rather than on
exhibits the highest values at γ=100, where
a large amount of shishes (~0.4 %) is formed, and the lowest values at γ=25, where less
shishes are formed. The nucleating ability of a substrate can be seen as the product between
the quality (efficiency) of the nucleation sites and their number. The quality of nucleation
sites is given by matching of the lattice parameters of the substrate and the crystallizing
polymer (epitaxy)38-41 and, in this case, is not an issue because of the common unit cell of
shishes and kebabs. In addition, it seems reasonable to assume that epitaxy matching does not
depend on the flow conditions. In contrast, the quantity of the nucleation sites is determined
by aspect ratio ( 1/D
) and concentration of the shishes and both depend on the flow
conditions. The individual effects are difficult to separate because a better nucleation ability
(higher values of T
) is associated with more nucleation sites that can be achieved both at
100
Chapter 5
the same concentration increasing the aspect ratio and with the same aspect ratio increasing
the concentration.
5.4.7 Melting of shish kebabs
Heterogeneous crystallization of kebabs on shishes yields a well organized and
oriented morphology. Because of the high nucleation temperature, kebabs grow thicker than
ordinary lamellae. For these reasons shish-kebabs exhibit an enhanced thermal stability. To
illustrate the melting of shish kebabs, Figure 5.13 shows SAXS intensities as a function of the
temperature during heating of a sample crystallized after a shear of 100 s
for 1 s at 142
.
Figure 5.13: SAXS equatorial and meridional integrated intensities as a function of
temperature heating a sample with fully oriented morphology crystallized on
cooling after a shear of 100
/1s at 142 .
At ~115
, melting of the kebabs (folded chain crystals) starts and it is completed
abruptly when they are superheated to ~142
. Shishes, with their extended chain structure,
are more stable than spherulites and melt gradually, surviving up to ~157
. Interestingly,
immediately after melting of the kebabs, the scattering fingerprint of shishes becomes again
visible in the SAXS. This indicates that, if the temperature is not too high, shishes generated
with shear at 142
retain their morphology after melting the kebabs. This behavior is
101
Chapter 5
similar to that of some nucleating agents. Kebabs that can melt leaving intact the shishes were
already observed and termed macrokebabs by Keller3. During cooling, after heating up to 148
, high temperature nucleation of kebabs is still observed, see Figure 5.14.
Figure 5.14: SAXS images during heating and cooling runs around the melting point of the
material.
This suggests that melting of the shishes begins with losing order at small length
scales. As a consequence, chain stretch rapidly vanishes and disrupts the structure of the
shishes. However, as suggested by some rheological observations42, the long range order, on
the length scale of the whole molecule, can still be retained. Apparently, the long lasting
orientation of HMW chains is sufficient for the epitaxial nucleation of kebabs. On heating
from room temperature, independently of the shear history at 142
~142
, kebabs always melt at
, i.e. a few degrees higher than spherulites do. In this respect, it is interesting the
follow the melting of a sample with a mixed morphology. For instance, Figure 5.15 reports
SAXS images of the melting of a sample crystallized after a shear of 5 s
102
for 10 s at 142
.
Chapter 5
Figure 5.15: Sequence of SAXS images describing morphological changes during melting of
a sample with a fully oriented morphology.
At low temperatures, this sample shows a morphology that is a combination of
oriented and isotropic distributed lamellae suggesting the coexistence of shish kebabs and
spherulites. Upon heating, the isotropic scattering coming from spherulites disappears first at
140
. The meridional lobes coming from kebabs disappear later, at 142
. Shishes were
not observed in SAXS during the crystallization of this sample and correspondingly are not
observed on heating after melting of the kebabs.
5.5 Conclusions
In this Chapter, we studied the mechanism of formation and melting of shish kebabs
in a sheared bimodal HDPE melt. The results indicate that flow induced precursors (FIPs)
play a crucial role in structural and morphological developments during crystallization. FIPs
are metastable structures with very low or no crystallinity that arise with stretching of the
longest chains in the melt. FIPs can be formed at temperatures surprisingly high compared
with T and, depending on the conditions, can grow into crystals or dissolve in the melt.
When FIPs are formed at very high temperatures, for instance 160 or 200
, they are
undetectable with SAXS and WAXD but still show strong influence on the crystallization of
the polymer. On cooling, FIPs increase the nucleation temperature and promote orientation at
structural and morphological levels. The formation of needle-like FIPs and shishes demands
stretch of the HMW chains. But this is not the only requirement. Shearing at 142
103
, we
Chapter 5
found that the total strain plays a clear role. It seems that a critical strain should be exceeded
to obtain a large quantity of shishes. On cooling, needle-like FIPs and shishes formed at 142
act as nucleating agents. They can raise the nucleation temperature to up 132
and
template the morphology of the polymer. With time resolved SAXS, it is possible to
differentiate between the high temperature epitaxial nucleation of kebabs and the nucleation
of lamellae randomly assembled in spherulites, at lower temperatures. We observe that ~0.4
% of shishes is sufficient to produce a fully oriented morphology on cooling. Such a
conclusion is consistent with our findings on iPP melts containing fibrils formed from the
nucleating agent DMDBS where a similar small amount (~0.5 wt%) of pre-aligned DMDBS
fibrils is sufficient to induce, on cooling, a fully oriented polymer morphology.
Shish kebabs obtained on cooling after shear at 142
are well organized and this
gives them an enhanced thermal stability. At morphological level, melting is a sequence of
events symmetrical to crystallization. In fully oriented morphologies, kebabs melt before
shishes and yield again SAXS patterns with equatorial scattering only. In mixed
morphologies, spherulites melt before kebabs and, similar to the crystallization process, no
equatorial scattering is observed.
In this Chapter, the transformation of FIPs into extended chain crystals and the
nucleating ability of these crystals is addressed. FIPs are partially disordered bundles of
molecules with undetectable crystallinity. In fast short term shear, a question arises whether
FIPs are generated during or after cessation of flow. In the following Chapter, an attempt has
been made to answer this question.
5.6 References
1.
2.
3.
4.
5.
Schrauwen, B. A. G.; Breemen, L. C. A. v.; Spoelstra, A. B.; Govaert, L. E.; Peters,
G. W. M.; Meijer, H. E. H. Macromolecules 2004, 37, 8618-8633.
Gahleitner, M.; Wolfschwenger, J.; Bachner, C.; Bernreitner, K.; Neiβl, W. Journal of
Applied Polymer Science 1996, 61, 649-657.
Keller, A.; Kolnaar, H. W. H., Flow induced orientation and structure formation.
VCH: New York, 1997; Vol. 18.
Meijer, H. E. H.; Govaert, L. E. Prog. Polym. Sci. 2005, 30, 915-938.
Ward, I. M.; Sweeney, J., An introduction to the mechanical properties of solid
polymers. John Wiley & Sons: Weinheim, 2004.
104
Chapter 5
6.
7.
8.
9.
10.
11.
12.
13.
14.
15.
16.
17.
18.
19.
20.
21.
22.
23.
24.
25.
26.
27.
28.
29.
30.
31.
32.
33.
34.
35.
Kumaraswamy, G.; Issaian, A. M.; Kornfield, J. A. Macromolecules 1999, 32, 75377547.
Janeschitz-Kriegl, H.; Ratajski, E.; Stadlbauer, M. Rheol Acta 2003, 42, 355.
Devaux, N.; Monasse, B.; Haudin, J. M.; Moldenaers, P.; Vermant, J. Rheol. Acta
2004, 43, 210-222.
Bashir, Z.; Odell, J. A.; Keller, A. Journal of Materials Science 1984, 19, 3713-3725.
Hill, M. J.; Barham, P. J.; Keller, A. Colloid & Polymer Sci. 1980, 258, 1023-1037.
Barham, P. J.; Keller, A. Journal of Materials Science 1985, 20, 2281-2302.
Hsiao, B. S.; Yang, L.; Somani, R. H.; Avila-Orta, C. A.; Zhu, L. Physical Review
Letters 2005, 94, 117802.
Smith, P.; Lemstra, P. J.; Pijpers, J. P. L.; Kiel , A. M. Colloid & Polymer Sci. 1981,
259, 1070-1080.
Hsieh, Y.; Ju, J. Journal of Applied Polymer Science 1994, 53, 347-354.
Loos, J.; Schimanski, T.; Hofman, J.; Peijs, T.; Lemstra, P. J. Polymer 2001, 42,
3827-3834.
Liu, T.; Lieberwirth, I.; Petermann, J. Macromol. Chem. Phys. 2001, 202, (14), 29212925.
Pennings, J.; Kiel , A. M. Kolloid ZZ Polym 1965, 205, 160.
Binsbergen, F. L. Nature 1966, 211, 516-517.
Hu, W.; Frenkel, D.; Mathot, V. B. F. Macromolecules 2002, 35, 7172-7174.
Dukovski, I.; Muthukumar, M. Journal of Chemical Physics 2003, 118, (14), 66486655.
Keum, J. K.; Burger, C.; Zuo, F.; Hsiao, B. S. Polymer 2007, 48, 4511-4519.
Zuo, F.; Keum, J. K.; Yang, L.; Somani, R. H.; Hsiao, B. S. Macromolecules 2006,
39, 2209-2218.
Kumaraswamy, G.; Kornfield, J. A.; Yeh, F.; Hsiao, B. S. Macromolecules 2002, 35,
1762-1769.
Ogino, Y.; Fukushima, H.; Matsuba, G.; Takahashi, N.; Nishida, K.; Kanaya, T.
Polymer 2006, 47, 5669-5677.
Keum, J. K.; Zuo, F.; Hsiao, B. S. Journal of Applied Crystallography 2007, 40, 4851.
Seki, M.; Thurman, D. W.; Oberhauser, J. P.; Kornfield, J. A. Macromolecules 2002,
35, 2583-2594.
Nogales, A.; Hsiao, B. S.; Somani, R. H.; Srinivas, S.; Tsou, A. H.; Balta-Calleja, F.
J.; Ezquerra, T. Polymer 2001, 42, 5247-5256.
Yang, L.; Somani, R. H.; Sics, I.; Hsiao, B. S.; Kolb, R.; Lohse, D. J. Phys. Condens.
Matter 2006, 18, 2421-2436.
Yang, L.; Somani, R. H.; Sics, I.; Hsiao, B. S.; Kolb, R.; Fruitwala, H.; Ong, C.
Macromolecules 2004, 37, 4845-4859.
Macosko, C., Rheology : principles, measurements, and applications. VCH:
Weinheim, 1994.
Balta-Calleja, F. J.; Vonk, C. G., X-ray scattering of synthetic polymers. Elsevier:
Amsterdam, 1989.
Wilchinsky, Z. W., Advances in X-ray analysis. 1962; Vol. 6.
Wunderlich, B., Thermal analysis of polymeric materials. Springer-Verlag: Berlin,
2005.
Kimata, S.; Sakurai, T.; Nozue, Y.; Kasahara, T.; Yamaguchi, N.; Karino, T.;
Shibayama, M.; Kornfield, J. A. Science 2007, 316, (5827), 1014 - 1017.
Coppola, S.; Balzano, L.; Gioffredi, E.; Maffettone, P. L.; Grizzuti, N. Polymer 2004,
45, 3249-3256.
105
Chapter 5
36.
37.
38.
39.
40.
41.
42.
Somani, R. H.; Yang, L.; Hsiao, B. S. Physica A 2002, 304, 145-157.
van Meerveld, J.; Peters, G. W. M.; Hutters, M. Rheol Acta 2004, 44, 119-134.
Wittmann, J. C.; Lotz, B. J. Polym. Sci., Polym. Phys. Ed. 1981, 19, 1837.
Thierry, A.; Straupe, C.; J., W.; Lotz, B. Macromol Symp 2006, 241, 103-110.
Fillon, B.; Thierry, A.; Lotz, B.; Wittmann, J. C. Journal of Thermal Analysis 1994,
42, 721-731.
Lotz, B.; Wittmann, J. C. J. Polym. Sci., Polym. Phys. Ed. 1986, 24, 1559.
Bent, J.; Hutchings, L. R.; Richards, R. W.; Gough, T.; Spares, R.; Coates, P. D.;
Grillo, I.; Harlen, O. G.; Read, D. J.; Graham, R. S.; Likhtman, A. E.; Groves, D. J.;
Nicholson, T. M.; McLeish, T. C. B. Science 2003, 19, 1691 - 1695.
106
Chapter 6*
Metastable structures during fast short term
shear
The goal of this Chapter is to study the crystallization of isotactic polypropylene
(iPP) during and immediately after a strong pulse of shear (i.e. with a high Deborah
number) on the undercooled melt. This is possible using high flux synchrotron X-ray
scattering (SAXS and WAXD) that allows for a high frame rate (3 frames/s)
maintaining a good signal to noise ratio. We found that shear rate is the dominating
parameter for structure formation during shear. When the shear rate is high enough,
crystals with a high degree of orientation in the shear direction can be formed already
during shear. In contrast, for lower shear rates, no crystalline structures were observed
during shear. However, SAXS shows an equatorial streak indicating needle-like
scatterers aligned with the flow direction. These scatterers are metastable precursors
of crystallization. Although these precursors are not crystalline in the early stages, they
crystallize after cessation of the flow and template the nucleation of lamellae with caxis parallel to the flow direction (shish-kebab morphology).
6.1 Introduction
Bulk crystallinity1-5 affects many of the physical properties of semicrystalline
polymeric materials6-12. Universal rules are difficult to establish as these materials exhibit a
processing-property relation, i.e. the final properties of the material are affected by the
processing conditions. Temperature history and, especially, flow alter the crystallization
behavior. Flow can increase the crystallization rate by decades compared to quiescent
conditions13-18 and, in addition, the crystalline morphology can change dramatically19-24. In
quiescent or quasi-quiescent conditions morphology is dominated by spherulites25, three
dimensional assemblies of randomly oriented folded chain lamellae. For strong enough flows,
spherulites are replaced by shish-kebabs, composite crystallites with an extended chain
fibrillar core (shish) dressed with disk-like folded chain lamellae (kebabs)26-32. The origin of
*
Partially reproduced from: L. Balzano et al. ‘Metastable structures during fast short term shear’,
Macromolecules 2008 (Submitted)
107
Chapter 6
this morphology is the topic of a long standing discussion31, 33-37 that begun in the 1960s. In
Chapters 4 and 5, some of the classic ideas are combined with new observations, to propose a
mechanism for the formation of shish-kebabs that starts with stretch of the longest chains and
passes through metastable precursors with undetectable crystallinity. In short term shear, after
cessation of the flow, a selection takes place among the precursors. Those exceeding some
critical dimension transform into crystalline structures, the remainder dissolve back into the
melt state. The ratio between crystallizing and dissolving precursors and the rate of
crystallization and dissolution are determined by the temperature. A real competition between
crystallization and dissolution of flow induced precursors (FIPs) can be observed only at high
enough temperatures. For instance, in the case of polyethylene (PE), it is necessary to be in
0
proximity to or even above the equilibrium melting point Tm0 ( Tm
PE
= 141.2 °C ). Although
there is no undercooling Δ T , this does not contradict thermodynamics. In fact, flow raises
the effective melting temperature by reducing the entropy and, thus, generates the
undercooling ΔT f that is indispensible for nucleation13, 38. At lower temperatures, the driving
force due to flow contribute to make most of the precursors stable from the moment of their
creation33. Here, dissolution is absent or limited to a small fraction of precursors.
In many of the experiments available in literature, the study of FIPs is based on short
term flow protocols33,
39-44
where, a pulse of flow is applied for few seconds, and the
evolution of the system is studied after cessation of the flow. In this scenario, some questions
arise: when are actually FIPs created and when do they crystallize? During or after cessation
of the flow? Reminding that FIPs are disordered bundles of molecules held together only by
weak interaction forces, another question that arises is: are these interaction forces strong
enough to prevent flow from destroying FIPs and allow them to crystallize during flow?
Giving answers to these questions is the aim of this Chapter.
In a short term flow protocol, flow lasts, typically, only for few seconds and this
restricts the sampling time to a fraction of a second. For this reason, our investigation
demands the use of synchrotron X-ray scattering with a high flux of photons. The material
investigated is isotactic polypropylene (iPP).
108
Chapter 6
6.2 Materials and methods
6.2.1 Materials
The polymer used in this work is a commercial homopolymer grade of isotactic
polypropylene (iPP). The material, labeled 15M10, was obtained by DSM (Geleen, The
Netherlands) and contained no other additives than stabilizers. The specifications of the
material are reported in Table 6.1.
Table 6.1: Specification of the iPP used in this work.
Material
Mw [kg/mol]
Mw/Mn
[%mmmm]
Tm [°C]
15M10
350
5.6
96.2
161
6.2.2 X‐ray characterization
Small angle X-ray scattering (SAXS) was performed at the beamline ID02 of the
European Synchrotron Radiation Facility (ESRF,Grenoble). The samples were irradiated with
a wavelength λ=0.995 Å and two dimensional images were recorded using a Frelon detector,
with a resolution of 1024x1024 pixels and a pixel size of 164 µm, placed at 6.5 m from the
sample. After subtraction of the scattering of the empty sample holder, images were
integrated to obtain the scattered intensity (I) as a function of the modulus of the scattering
vector q = ( 4π / λ ) sin(θ / 2) where 2θ is the scattering angle45, 46. The total intensity I SAXS
as a function of time was obtained integrating I(q) over the whole accessible q range:
I
SAXS
(t ) =
qmax
∫
I (q; t )dq .
qmin
Wide angle X-ray scattering (WAXS or WAXD) was performed at the beamline ID11
of the ESRF. Experiments were carried out with a wavelength λ=0.508 Å and a sample to
detector distance of about 32cm. The images were recorded with a two dimensional Frelon
detector with a resolution of 512x512 pixels (pixel size 190 µm). After correction for spatial
109
Chapter 6
distortions and scattering of the empty sample holder, the images were integrated to obtain
the intensity distribution I as a function of the scattering angle 2θ . These one-dimensional
profiles were used to calculate crystallinity. For this purpose, the scattering of the amorphous
component (IA) underneath crystalline peaks (Ic) was approximated with a straight line (see
Figure 6.1).
Figure 6.1: Example of the procedure to separate crystalline and amorphous scattering. The
Miller indices of the scattering reflections are indicated as well.
This simplified procedure is expected to give trustworthy results at low crystallinities
(up to ~15%) 47. Crystallinity (x) was determined as:
x = 100 ⋅
IC
IC + I A
(6.1)
WAXD data were also used for calculating the Hermans’ orientation factors FH of the
crystalline c-axis:
FH =
3 cos 2 β − 1
2
110
(6.2)
Chapter 6
where β is the azimuthal angle. For iPP, where a 00l reflection is not observed, cos 2 β can
be obtained with Wilchinsky’s formula48. Using geometrical relations, the orientation of the
c-axis is calculated from the orientation of the 110 and 040 crystal planes:
cos 2 β = 1 − 0.901 cos 2 β 040 − 1.099 cos 2 β110
(6.3)
with the definition:
π /2
cos β =
2
∫ I (β ) ⋅ cos
0
2
β ⋅ sin β ⋅ d β
π /2
(6.4)
∫ I (β ) ⋅ sin β ⋅ d β
0
6.2.3 Shear experiments
Shear experiments were performed in combination with SAXS and WAXD using a
Linkam CSS-450 Shear Cell where, to avoid unwanted scattering, the original glass plates
were replaced with kapton. The setup is shown in Figure 6.2.
Figure 6.2: Schematic drawing of the shear device.
When shearing, a metal spoke of the rotating plate lies periodically in front of the
incoming beam, the corresponding (dark) scattering images are discarded from analysis.
111
Chapter 6
6.3 Results and Discussion
6.3.1 Flow conditions in short term shear
It is well established that shear flow promotes orientation and stretch of polymer
molecules in the melt23,
24, 49-51
. If τ D and τS are the relaxation times for orientation
(disengagement time) and stretch (chain retraction time) of the longest molecules, orientation
of these molecules is achieved when the Deborah number for orientation DeO = τ D γ
exceeds unity and stretch when the Deborah number for stretch DeS = τS γ exceeds unity49.
The ratio τ D / τ S is similar to the number of entanglements per chain Z
52-54
. For the iPP
under consideration, Z 100 and, therefore, DeO ≅ 100DeS . In other words, chain orientation
is attained at shear rates nearly two decades lower than chain stretch. In this Chapter, we aim
to look at crystallization under shear conditions providing orientation and stretch to the
molecules, therefore, our experimental conditions are chosen such that DeS >>1. A high
DeO is consequently achieved as well.
A useful way to study the flow induced crystallization (FIC) of a polymer melt is
short term shear. In this experiment, after annealing at high temperature, the melt is brought
to the desired test temperature where shear is applied for a limited time, typically few
seconds, and the subsequent crystallization of the polymer is observed in absence of flow. By
choosing a short shear time it is hoped that flow effects are separated from crystallization
effects, i.e. no noticeable material changes occur during the flow and thus the ‘normal’
rheological behavior can be used to characterize the flow strength ( DeO and DeS ). The flow
conditions are determined by three parameters: temperature, shear rate ( γ ) and shear time ( ts
). Once these are assigned, the total strain ( γ = γ ⋅ ts ) is also fixed. A fair comparison
between different experiments is possible only when holding γ constant.
112
Chapter 6
6.3.2 Flow induced precursors during short term shear
To generate FIPs, we select shear conditions with a relatively high strain value,
γ = 180 . This allows for spanning a broad range flow conditions, from relatively slow and
long to relatively fast and short. FIPs originate from stretching of the longest chains in the
melt
23, 28, 41, 44, 55
, therefore we restrict γ such as DeS >> 1 . For the iPP considered,
τ D ≅ 100 s and τ S ≅ 1 s at 145 °C
56
. Therefore, shear rates of 60 s-1, 90 s-1 and 180 s-1
suite our experimental requirements. The shear time is varied accordingly to keep the total
strain γ = 180 . Table 6.2 gives an overview of the selected shear conditions. WAXD patterns
acquired at a rate of 3 frames/s during and immediately after short term shear at 145 °C are
shown in Figure 6.4. In these figures, as in the rest of the Chapter, the time t=0 corresponds
to cessation of the flow. During flow (t 1 ). As revealed by the appearance of SAXS meridional lobes, after cessation of the
flow, FIPs start crystallizing and serve as a heterogeneous substrate for the nucleation of
stack of lamellae with c-axis parallel to the flow direction.
Figure 6.3: SAXS patterns during and immediately after short term shear at 145 °C with 60s1
/3s.
6.3.3 Crystallization after short term shear.
After cessation of shear (t=0), samples are allowed to crystallize isothermally at 145
SAXS
SAXS
°C. Figure 6.5 reports the SAXS intensities in the equatorial ( I Eq ) and meridional ( I Mer )
114
Chapter 6
Figure 6.4: Equatorial region of WAXD patterns during and after short term shear at 145°C
with γ =180. Flow direction is vertical. The arrows indicate arched crystalline
reflections.
115
Chapter 6
SAXS
SAXS
regions for the experiment 60s-1/3s. At the beginning, I Eq
is much higher than I Mer
SAXS
because of the equatorial streak coming from needle like FIPs. However, while I Eq
SAXS
remains nearly constant, I Mer grows quickly and after few seconds the situation is reversed,
SAXS
SAXS
SAXS
i.e. I Mer > I Eq . The growth of I Mer is a consequence of the onset of meridional lobes
associated with the decoration of shishes and FIPs with stacks of lamellae (kebabs) with the
c-axis oriented with the flow direction. The formation of a similar shish-kebab morphology is
observed also for 90s-1/2s and 180s-1/1s after cessation of flow.
Figure 6.5: SAXS meridional and equatorial intensities at 145 °C as a function of time after
the application of shear (60s-1/3s).
As shown in Figure 6.6, the growth of kebabs raises the crystallinity and,
simultaneously, the crystalline orientation factors (Figure 6.7) tend to decrease. This can be
explained with the homogeneous nucleation of randomly oriented lamellae in the bulk of the
sample and/or with bending of the kebabs after a certain diameter is exceeded (see Figure
6.8)61. In the experiments 90s-1/2s and 180s-1/1s, where crystallization is observed already
during shear, the orientation is significantly higher than in 60s-1/3s. As expected, the data of
Figure 6.6 indicate that an increase in γ is more efficient to enhance both nucleation and
orientation than an increase in ts .
116
Chapter 6
Figure 6.6: Crystallinity developing at 145 °C after the application of 180 strain units.
Figure 6.7: Orientation developing at 145 °C after the application of 180 strain units.
117
Chapter 6
Figure 6.8: Bending of kebabs occurring after a certain diameter is exceeded.
More details on structure formation immediately after cessation of shear can be
obtained defining a degree of space filling after flow:
Φ(t ) =
x(t ) − x0
x∞ − x0
(6.5)
x0 is the crystallinity already present at the time t=0 and, therefore, generated during flow,
while, x∞ is the saturation level for the crystallinity, attained at long times. Note that the
definition (6.5) is based on the assumption that, at long times, a complete space filling ( Φ =1)
is attained despite the relatively low degree of crystallinity ( x∞ ). Φ(t ) can be described with
the Avrami equation62-64:
Φ(t ) = 1 − exp(−kt n )
(6.6)
where k is a rate constant and n the Avrami exponent that indicates the geometry of the
growing crystallites. Equation (6.6) can be re-written as:
Ln {− Ln [1 − Φ (t ) ]} = Ln ( k ) + n ⋅ Ln (t )
(6.7)
In the early stages, Ln {− Ln [1 − Φ (t ) ]} as a function of Ln (t ) (the so called Avrami
plot) is a straight line with slope n. The Avrami plot, for the crystallization following the
cessation of the shear, is shown in Figure 6.9. Fitting Equation (6.7) on the data yields the
parameters given in Table 6.3.
118
Chapter 6
Table 6.3: Parameters used for fitting Equation (6.6) onto crystallization data of Figure 6.6.
x∞ [%]
x0 [%]
n
k [s-n]
60 s-1/3s
14.5
0
1.34
2.70·10-3
90 s-1/2s
16.2
0.1
1.37
6.78·10-3
180 s-1/1s
18.8
2
1.38
8.20·10-3
Figure 6.9: Avrami plot constructed with crystallinity data after short term shear at 145 °C.
Independent of the flow conditions, n 1.36±0.02 indicates that, in all cases, space
filling occurs because of the athermal growth of circular lamellae (kebabs). The only
difference is the rate constant k that increases with γ , in line with the increased number of
kebabs at higher shear rates and consistent with an increased number of shishes as well.
Combining equations (6.5) and (6.6), it is possible to obtain a description of the crystallinity
as a function of time:
x(t ) = x0 + ( x∞ − x0 ) ⎡1 − exp ( −kt n ) ⎤
⎣
⎦
(6.8)
As expected, using the fitting parameters given in Table 6.3, a good agreement with
the measured data is found (lines in Figure 6.6). Some remarkable observations emerge from
this analysis:
119
Chapter 6
•
After a short term shear, a good description of the crystallization kinetics can
be obtained with the Avrami equation. To this end, it is necessary to define the
space filling Φ(t ) accounting for the crystallinity generated already during
flow ( x0 in Equation (6.5)). Although this value can be very small, it has large
influence in the early stages. Furthermore, a complete space filling at long
times needs to be hypothesized, despite the relatively low crystallinity.
•
The data of Figure 6.6 seem to indicate the presence of an ‘induction time’ for
the growth of kebabs, even for 180s-1/1s where a crystallinity of 2 % is
attained already during flow. However, although in all experiments there is a
clear characteristic time at which x starts increasing massively, this induction
period should not be considered as a time where structure formation is frozenin. Figure 6.9 indicates that, in all cases, starting form the first measurement
(t=0.3 s), major changes take place in the space filling.
•
At long times, in isothermal conditions, crystallinity saturates to a plateau
level ( x∞ ) that depends on the flow conditions. Higher γ yield higher final
crystallinity. This behavior can be interpreted in terms of driving force for
crystallization, i.e. in terms of Gibbs free energy (G) difference between the
melt and the crystal. Flow increases the free energy of the melt generating
more driving force for nucleation (see Figure 6.10).
Figure 6.10: Effect of flow on the free energy.
120
Chapter 6
6.4 Conclusions
We studied the early stages of the flow induced crystallization of iPP, during and
immediately after the application of a strong (high De) shear pulse. At 145 °C, with a total
strain γ = 180 , X-ray scattering data (SAXS and WAXD) suggest that shear rate is the
dominant parameter for structure development during flow. When the shear rate is high
enough (> 90 s-1) crystalline structures can be formed already during shear. In these
conditions, WAXD suggests the presence of crystallinities up to 2% and a high degree of
orientation of the crystals with the flow direction. In contrast, at lower shear rates (for
instance 60 s-1) no sign of crystalline structures is observed during shear. However,
simultaneously, SAXS indicates the presence of needle like scatterers aligned with the shear
direction. As soon as the flow stops, the crystalline or non-crystalline structures generated
during shear act as a nucleating substrate for the heterogeneous formation of kebabs. The
crystalline morphology is, thus, strongly tied with the structures present in the early stages.
6.5 References
1.
2.
3.
4.
5.
6.
7.
8.
9.
10.
11.
12.
13.
14.
15.
Strobl, G., The physics of polymers. Springer: Berlin, 1997.
Wunderlich, B., Thermal analysis of polymeric materials. Springer-Verlag: Berlin,
2005.
Keller, A. Phyl. Mag. 1957, 2, 1171.
Lotz, B. Journal of Macromolecular Science, Part B-Physics 2002, B41, 685-709.
Natta, G. J. Polym. Sci. 1955, 16, 143.
Schrauwen, B. A. G.; Breemen, L. C. A. v.; Spoelstra, A. B.; Govaert, L. E.; Peters,
G. W. M.; Meijer, H. E. H. Macromolecules 2004, 37, 8618-8633.
Kristiansen, M.; Werner, M.; Tervoort, T.; Smith, P.; Blomehofer, M.; Schmidt, H.
W. Macromolecules 2003, (36), 5150-5156.
Smith, P.; Lemstra, P. J. Journal of Materials Science 1980, 15, 505-514.
Baastiansen, C. W. M., Oriented structures based on flexible polymers. PhD thesis,
Eindhoven University of Technology: 1991.
Govaert, L. E., Deformation behavior of oriented polyethylene fibers. PhD thesis,
Eindhoven University of Technology: 1990.
Meer, D. W. v. d.; Pukanszky, B.; Vancso, G. J. J. Macromol. Sci.-Phys. 2002, 41,
1105-1119.
Bashir, Z.; Odell, J. A.; Keller, A. Journal of Materials Science 1984, 19, 3713-3725.
Coppola, S.; Balzano, L.; Gioffredi, E.; Maffettone, P. L.; Grizzuti, N. Polymer 2004,
45, 3249-3256.
Lagasse, R. R.; Maxwell, B. Polym. Eng. Sci. 1976, 16, 189.
Heeley, E. L.; Morgovan, A.; Bras, W.; Dolbnya, I. P.; Gleeson, A. J.; Ryan, A. J.
Phys. Chem. Comm. 2002, 5, 158-160.
121
Chapter 6
16.
17.
18.
19.
20.
21.
22.
23.
24.
25.
26.
27.
28.
29.
30.
31.
32.
33.
34.
35.
36.
37.
38.
39.
40.
41.
42.
43.
44.
45.
Kumaraswamy, G.; Kornfield, J. A.; Yeh, F.; Hsiao, B. S. Macromolecules 2002, 35,
1762-1769.
Nogales, A.; Hsiao, B. S.; Somani, R. H.; Srinivas, S.; Tsou, A. H.; Balta-Calleja, F.
J.; Ezquerra, T. Polymer 2001, 42, 5247-5256.
Devaux, N.; Monasse, B.; Haudin, J. M.; Moldenaers, P.; Vermant, J. Rheol. Acta
2004, 43, 210-222.
Keller, A.; Kolnaar, H. W. H., Flow induced orientation and structure formation.
VCH: New York, 1997; Vol. 18.
Hsiao, B. S.; Yang, L.; Somani, R. H.; Avila-Orta, C. A.; Zhu, L. Physical Review
Letters 2005, 94, 117802.
Hobbs, J. K.; Humphris, A. D. L.; Miles, M. J. Macromolecules 2001, 34, 5508-5519.
Sakellarides, S. L.; McHugh, A. J. Rheol Acta 1987, 26, 64-77.
Seki, M.; Thurman, D. W.; Oberhauser, J. P.; Kornfield, J. A. Macromolecules 2002,
35, 2583-2594.
Somani, R. H.; Hsiao, B. S.; Nogales, A.; Srinivas, S.; Tsou, A. H.; Sics, I.; BaltaCalleja, F. J.; Ezquerra, T. A. Macromolecules 2000, 33, 9385-9394.
Basset, D. C.; Franck, F. C.; Keller, A. Phyl Trans Roy Soc London A 1994, 348, 2943.
Binsbergen, F. L. Nature 1966, 211, 516-517.
Pennings, J.; Kiel , A. M. Kolloid ZZ Polym 1965, 205, 160.
Ogino, Y.; Fukushima, H.; Matsuba, G.; Takahashi, N.; Nishida, K.; Kanaya, T.
Polymer 2006, 47, 5669-5677.
Wang, Z. G.; Hsiao, B. S.; Sirota, E. B.; Srinivas, S. Polymer 2000, 41, 8825-8832.
Yamazaki, S.; Watanabe, K.; Okada, K.; Yamada, K.; Tagashira, K.; Toda, A.;
Hikosaka, M. Polymer 2005, 46, 1685-1692.
Kimata, S.; Sakurai, T.; Nozue, Y.; Kasahara, T.; Yamaguchi, N.; Karino, T.;
Shibayama, M.; Kornfield, J. A. Science 2007, 316, (5827), 1014 - 1017.
Hill, M. J.; Barham, P. J.; Keller, A. Colloid & Polymer Sci. 1980, 258, 1023-1037.
Janeschitz-Kriegl, H.; Eder, G. Journal of Macromolecular Science, Part B 2007, 46,
591-601.
Kanaya, T.; Takayama, Y.; Ogino, Y.; Matsuba, G.; Nishida, K. Lect. Notes Phys.
2007, 714, 87-96.
McHugh, A. J. Polym. Eng. Sci. 1982, 22, (15-26).
Somani, R. H.; Yang, L.; Zhu, L.; Hsiao, B. S. Polymer 2005, 46, 8587-8623.
Yamazaki, S.; Hikosaka, M.; Toda, A.; Wataoka, I.; Gu, F. Polymer 2002, 43, 65856593.
Coppola, S.; Grizzuti, N.; Maffettone, P. L. Macromolecules 2001, 34, (14), 5030 5036.
Azzurri, F.; Alfonso, G. C. Macromolecules 2005, 38, 1723-1728.
Keum, J. K.; Zuo, F.; Hsiao, B. S. Journal of Applied Crystallography 2007, 40, 4851.
Jerschow, P.; Janeschitz-Kriegl, H. International Polymer Processing 1997, 12, (1),
72-77.
Van der Beek, M. H. E.; Peters, G. W. M.; Meijer, H. E. H. Macromolecules 2006,
39, (5), 1805 -1814.
Pogodina, N. V.; Winter, H. H.; Srinivas, S. Journal of Polymer Science: Part B:
Polymer Physics 1999, 37, 3512-3519.
Elmoumni, A.; Gonzalez-Ruiz, R. A.; Coughlin, E. B.; Winter, H. H. Journal of
Chemical Physics 2005, 206, 125-134.
Stribeck, N., X-ray scattering of soft matter. Springer Laboratory: 2007.
122
Chapter 6
46.
47.
48.
49.
50.
51.
52.
53.
54.
55.
56.
57.
58.
59.
60.
61.
62.
63.
64.
Balta-Calleja, F. J.; Vonk, C. G., X-ray scattering of synthetic polymers. Elsevier:
Amsterdam, 1989.
Kumaraswamy, G.; Verma, R. K.; Kornfield, J. A.; Yeh, F.; Hsiao, B. S.
Macromolecules 2004, 37, (24), 9005 -9017.
Wilchinsky, Z. W., Advances in X-ray analysis. 1962; Vol. 6.
van Meerveld, J.; Peters, G. W. M.; Hutters, M. Rheol Acta 2004, 44, 119-134.
Dukovski, I.; Muthukumar, M. Journal of Chemical Physics 2003, 118, (14), 66486655.
Somani, R. H.; Yang, L.; Hsiao, B. S. Physica A 2002, 304, 145-157.
Dealy, J. M.; Larson, R. G., Structure and Rheology of Molten Polymers. Hanser
Gardner Pubns: Cincinnati, 2006.
Doi, M.; Edwards, S. F., The theory of polymer dynamics. Clarendon Press: Oxford,
1986.
McLeish, T. C. B. Advances in Physics 2002, 51, (6), 1379-1527.
Yang, L.; Somani, R. H.; Sics, I.; Hsiao, B. S.; Kolb, R.; Lohse, D. J. Phys. Condens.
Matter 2006, 18, 2421-2436.
Peters, G. W. M.; Swartjes, F. H. M.; Meijer, H. E. H. Macromol. Symp. 2002, 185,
277-292.
Fujiyama, M.; Wakino, T. Journal of Applied Polymer Science 1988, 35, 29-49.
Kolb, R.; Seifert, S.; Stribeck, N.; Zachmann, H. G. Polymer 2000, 41, 1497-1505.
Ran, S.; Zong, X.; Fang, D.; Hsiao, B. S.; Chu, B. Macromolecules 2001, 34, 25692578.
Ruland, W. Journal of Polymer Science: Part C 1969, 28, 143-151.
Xu, J.; Johnson, M.; Wilkes, G. L. Polymer 2004, 45, (15), 5327-5340.
Avrami, M. J. Chem. Phys. 1939, 7, 1103-1112.
Avrami, M. J. Chem. Phys. 1940, 8, 212-224.
Avrami, M. J. Chem. Phys. 1941, 9, 177-184.
123
Chapter 6
124
Chapter 7
Conclusions and recommendations
7.1 Conclusions / Technology assessment
Polyolefins, basically polyethylene (PE) and polypropylenes (PP), are materials with a
vast number of applications. However, nearly seventy years after the industrial launch, their
potential is not yet fully exploited. Smart processing, additives and control over molecular
parameters can still enhance the performances of these olefin-based materials.
The mechanical, optical and transport properties of polyolefins are strictly related to
crystallinity and crystalline morphology. For this reason, one of the key issues is
understanding the crystallization behavior, especially in presence of flow, in order to be able
to direct it in the desired way. Flow can enhance the crystallization kinetics, can change the
morphology and can make crystallization happen even at temperatures where it is not
expected, i.e. above the melting point (extended chain crystals).
During processing, the flow conditions (geometry, temperature, flow rate) should be
tailored in primis to the molecular weight and to the molecular weight distribution of the melt
but also to the additives that are present.
Self-assembling sorbitol-based additives (DBS, MDBS, DMDBS, etc.) are often used
as nucleating/clarifying agent for isotactic polypropylene (iPP). At concentrations between
0.1 and 1 wt%, these additives can crystallize before iPP, leading to an interconnected
network of nano-fibrils. The polymer crystallizes forming disk-like lamellae that grow in the
direction of the radius of these nanofibrils. As a consequence, the orientation of iPP lamellae
is controlled by the orientation of the nanofibrils at the onset of iPP crystallization. In absence
of flow, DMDBS nanofibrils grow randomly oriented in the melt but they can be aligned with
flow. Like in many other cases, flow can switch lamellar orientation on and off. However,
when sorbitol based additives are used, there is a rather sharp transition between the flow
conditions that can yield orientation and those that cannot. This transition is the
crystallization of the additive. When a random lamellar orientation is desired, flow should be
125
Chapter 7
applied at temperatures higher than the crystallization of the additive. On the other hand,
when a high degree of lamellar orientation is desired, flow should be applied at temperatures
lower than the crystallization of the additive. Such a procedure, in combination with ~0.5
wt% of DMDBS, can lead to a fully oriented iPP morphology after the application of flow at
a temperature as high as ~200 °C.
Heterogeneous nucleation on oriented substrates can be used in several processes (e.g.
spinning) where a high degree of lamellar orientation in the final product is demanded. The
outstanding performances of sorbitol-based additives in iPP are due to: 1) a good dispersion
in the polymer matrix; 2) a high aspect ratio of the nanofibrils; 3) a good epitaxy matching.
Similar conditions, with a pre-aligned nano-fibrillar substrate templating the morphology of
the bulk of the polymer, can be generated also in polyethylene (PE). Here the nanofibrils are
made out of the same polymer and can be produced in-situ by inducing, with flow, the
extended chain crystallization of the longest molecules of the melt. The molecular weight
distribution of the melt needs to be carefully tailored on the flow conditions. Very broad and
bimodal molecular weight distributions can fulfill the requirements as demonstrated in the
thesis.
7.2 Recommendations for future research
Like all scientific contributions, this thesis tries to give few answers and puts forward
many questions. Several aspects demand further clarification. Two outstanding questions are:
• Heterogeneous crystallization is based on the epitaxial relation (lattice
matching) between the nucleating agent and the crystallizing polymer.
Polymer molecules prefer to crystallize onto heterogeneous substrates
because, below the melting point, the contact (adsorption) of a polymer
molecule onto the foreign particle minimizes the energy of the system. What
happens above the melting point? Is this scenario retained? And what are the
consequences on the rheology of the melt?
• Formation of needle-like flow induced precursors with undetectable
crystallinity is a typical example of ‘SAXS before WAXD’. Flow induced
precursors assume the form of structures with only short range order. In this
126
Chapter 7
thesis, by calling them precursors, we confine them in a black box. What is
the structure of flow induced precursors? Do they differ from the precursors in
quiescent conditions? Do they alter the rheology of the system? Do chains
need to disentangle to generate a FIP?
• Polymers with broad molar mass distributions are often used in applications
(e.g. film blowing). In these melts, there is a subtle balance between the
relaxation times of long and short chains based on the life time of
entanglements that can be very limited (constraint release). For the formation
of oriented morphologies, what is more important? The high or the low molar
mass?
127
Chapter 7
128
Samenvatting
Synthetische polymeren worden momenteel geproduceerd in grote hoeveelheden,
meer dan 200 miljoen ton op jaarbasis. Meer dan 70% van dit produktie volume wordt
ingenomen door de zogenaamde klasse van “Commodity Plastics”, te weten de bekende vier
bulkpolymeren:
polyethyleen
(PE),
polypropyleen
(PP),
polystyreen
(PS)
en
polyvinylchloride (PVC). Polyethyleen en Polypropyleen vormen samen de subgroep
Polyolefinen en deze klasse van polymeren is met meer dan 100 miljoen ton op jaarbasis de
belangrijkste groep van industriële polymeren.
De aanduiding Polyolefinen of PE en PP is generiek, bijv. er worden van PE en PP tal
van soorten geproduceerd die onderling verschillen in molmassa (ketenlengte), ketenstructuur
(stereoregulariteit, vertakkingen etc.), allemaal toegesneden op specifieke toepassingen.
De eigenschappen van synthetische polymeren worden niet alleen bepaald door de
chemische structuur van het polymeer maar ook door de verwerkingscondities tot eind
produkt. Ketenoriëntatie is met name bepalend voor de mechanische eigenschappen met als
extreem voorbeeld het verschil in eigenschappen tussen flexibele folies en containers en de
supersterke vezel Dyneema® van DSM, allemaal gebaseerd op (lineair) PE maar met een
verschil
in
keten
oriëntatie
en
molmassa.
Naast
de
chemische
structuur
en
verwerkingscondities kan men de eigenschappen van polymeren ook sterk beïnvloeden door
het gebruik van additieven en er zijn tal van additieven ontwikkeld zoals kleurstoffen,
brandvertragers, anti-oxydanten en om bijv. de vloei van de polymeren te verbeteren bij
verwerking of ter verbetering van de mechanische eigenschappen zoals glasvezels ter
verhoging van de stijfheid. In feite is een plastic (compound) een polymeer + additief!
In dit proefschrift is aandacht besteed aan het phenomeen van nucleatie, het begin van
het kristallisatie-proces. Polymeren zoals PE en PP kunnen kristalliseren en het optreden van
kristallisatie is zeer belangrijk voor de mechanische eigenschappen. Zonder kristallisatie
zouden polymeren zoals PE en PP rond kamertemperatuur en/of verhoogde temperatuur niet
gebruikt kunnen worden omdat het dan slechts hoog-visceuze vloeistoffen zouden zijn.
Kristallisatie c.q. ordening van de lange polymeer moleculen in de bekende zogenaamde
gevouwen-keten kristallen (lamellar folded-chain crystals) geeft versteviging aan PE en PP
129
tot aan het smeltpunt. Polypropeen (PP) is een uiterst traag kristalliserend polymeer en dat is
inherent aan de structuur van het molecuul, een 31 helix in de kristallijne fase. Bij
vormgeving, bijv. spuitgieten, is dit kristallisatie proces te traag. Om de overall
kristallisatiesnelheid te verhogen worden zogenaamde kiemvormers toegevoegd die het
ontstaan van kiemen, resulterend in kristallen, in het materiaal sterk bevorderen.
Er zijn twee typen kiemvormers bestudeerd, te weten:
-
De bekende sorbitol derivaten die commercieel worden gebruikt en die als
eigenschap hebben dat ze, gemengd met PP, assembleren in nano-fibrillen bij
afkoelen voor dat de PP fase kristalliseert. Op deze nano-fibrillen kan PP
gemakkelijk ontkiemen;
-
Een mengsel van hoog- en laag-moleculair PE, gemaakt via een unieke
zogenaamde “one pot” synthese waardoor de beide componenten op moleculaire
schaal zijn gemengd. Onder afschuiving, zoals optreedt bij verwerking via de
polymere smelt, zal de hoog-moleculaire component gemakkelijker orienteren en
deze gestrekte ketens fungeren als kiem voor kristallisatie van de laag-moleculaire
matrix PE.
Bij het PP/sorbitol systeem werd gevonden dat bij zeer lage concentraties, < 1%, van het
gebruikte type 1,3:2,4-bis (3,4-dimethylbenzylideen) sorbitol (DMDBS) zeer efficiënte
kiemvorming optreedt en resulteert in zeer kleine PP kristallieten die nagenoeg geen licht
verstrooien, dus geeft een transparant PP produkt. DMDBS is oplosbaar in gesmolten PP
maar kristalliseert uit in nano-fibrillen, door zelf-assemblage, die fungeren als kiem voor
kristallisatie van PP. Door gebruik te maken van gecombineerde temperatuur en
afschuifprofielen kunnen de DMDBS nano-fibrillen worden georiënteerd in een bepaalde
richting en dit biedt de mogelijkheid om unieke structuren te maken.
Bij het PE/PE systeem, respectievelijk laag- en hoog-moleculair PE, werd gevonden dat de
hoog-moleculaire component kan worden georiënteerd boven het smeltpunt van de matrix,
de laag-moleculaire component. Dit biedt de mogelijkheid tot nucleatie van de laagmoleculaire component op de gestrekte ketens van hoog-moleculair PE., in feite
kristalliseert de laag-moleculaire PE component op hetzelfde materiaal, een perfecte vorm
van epitaxy!
130
In conclusie, er werden twee systemen bestudeerd die ogenschijnlijk niets met elkaar te
maken hebben maar het samenbindend element is het genereren van georiënteerde kiemen
resulterend in unieke structuren c.s. morphologie van respectievelijk PP en PE.
131
132
Acknowledgements
Many people have contributed to this thesis and many have contributed to make my
last four years in Eindhoven a real good time, I want to express my gratitude to everyone.
First of all, I would like to acknowledge the people who offered me the opportunity of
doing a PhD and provided the necessary resources: Prof. Piet Lemstra, Prof. Sanjay Rastogi
and Dr. Gerrit Peters.
I am grateful to Prof. Piet Lemstra for the motivations, the teachings and for instilling
in my mind his research style, deeply scientific but always with an eye to applications. This
work could not be done without Prof. Sanjay Rastogi, a daily source of inspiration. Thank
you, Sanjay, for all the discussions and for the stimulating ideas that led to this thesis.
Working with you, I learned a lot, challenging or validating the existing ideas, always with
determination and enthusiasm. In addition, I want to thank you for the friendship that made
me always feel comfortable. This thesis is largely due to Dr. Gerrit Peters. Hartelijk bedankt,
Gerrit, for your inputs (including the idea of FIP) and for examining thoroughly the results
with constructive criticism. You thought me to read the physics beyond the data and showed
me positive thinking and great competence.
For this thesis, I made large use of X-rays. In my 22 (!) synchrotron (ESRF)
experiments, I benefitted of the help of many people. I am grateful to the staff of the
beamlines BM26/ DUBBLE, ID02 and ID11 of the ESRF for supporting our experiments.
Here, I would like to mention the persons that introduced me to X-ray scattering: Dr. Guido
Heunen and Dr. Ann Terry. Guido cleverly arranged setups at BM26 at all times of the day
and of the night. Ann is a tireless coworker that solved many setup and software deadlocks at
ID02 and ID11. The discussions with Wim Bras were important in building motivations. I
also acknowledge Dr. Giuseppe Portale for the support at BM26, Dr. Caroline Curfs for ID11
and of Dr. Peter Boesecke for ID02. Colleagues from SKT and M@te gave their valuable
contribution to several ESRF experiments. Above all, I am indebted with Jan-Willem
Housmans for all the time spent at the ESRF, including the always present dropjes.
I had good time working in SKT where I benefitted of the support of several people. I
would like to thank Dr. Han Goossens for the discussions and the comments, especially in the
early stages of my PhD and Dr. Juan Fran Vega for teaching me the tricks of rheometry. My
gratitude extends to Dirk Lippits for the stimulating discussions about rheology and for
133
calling me ‘maestro’, to Jules Harings for thorough discussions about my and his research
that helped in building motivations, to Esther Vinken for the help with TGA and ‘that’ sms,
to Nilesh Kukalyekar for the fruitful collaboration, to Irina Cotiuga for the discussions and
the good time, to Joachim Loos, Anne Spoelstra and Pauline Schmit for microscopy, to
Denka Hristova for the help in solving my initial problems with X-rays, to Sachin Jain for the
discussions, to Joost Valeton for DSC, to Jules Kierkels for the help with X-rays and to
Sainath Vaidya for the CNTs solutions. The list includes also my officemates at STO 0.42
Mano Prusty, Dirk Lippits and Gizela Mikova and the other people of SKT and PTG: Laurent
Nelissen, Peter Koets, Mark van der Mee, Chunxia Sun, Lijing Xue, Rafiq Ahmed, Marjolein
Diepens, Marjoleine Drieskens, Roy l’Abee, Saeid Talebi, Weizhen Li, Maya Ziari, Bjorn
Teurlings, Edgar Karssenberg, Soney Varghese, Martijn Jansen, Cees Weijers, Thierry
Leblanc. Thanks to Bob Fifield for all the incursions in my office, the handshakes and the
always open debate about the reason why the Romans did not invade Wales. Thanks to Elly
Langstad and Ineke Kollenburg for all the care. From the group M@te, I am grateful to
Reinhard Forstner, Roel Janssen, Frederico Custodio and Rudi Steenbakkers for the
discussions and the good time together, also at the ESRF. I also want to thank Sjef Garenfeld
for providing help to my last minute requests.
My life in Eindhoven was always hectic thanks to some memorable characters. I was
introduced en la movida de Eindhoven by Carlos. Graciaaaas, Carlos, for being a real friend.
All the events (including the FORT) together with Blanca, David, Ben and Chiara have
something to be remembered. I enjoyed these times like there is no tomorrow. Unforgettable.
This list cannot be closed without acknowledging the friendship of Dillip (is he spanish?), the
class of Brett, the black clothing of Chris, the bbqs of Dani, the world according to Kuba, the
dinners of Lorenzo, the sweetness of Mari, the twisted reasoning of Vincente, the French
accent of Emilie, the smiling eyes of Amparo, the sincerity of Víctor and the smiles of
Monica.
Last and certainly not least, Chiara. I take the risk of being banal saying that you gave
me incredible support. We had great time and you also tolerated my difficult moments. I
thank you for the comprehension and the patience.
Per chiudere, ringrazio la mia famiglia per tutto il supporto e gli incoraggiamenti
ricevuti e quella Rabotti per la stima e le vacanze in Toscana.
134
Curriculum Vitae
The author was born in Pompei, Italy, on June 25th 1977. After finishing secondary
school (Liceo Scientifico ‘E.Pascal’, Pompei), he studied Chemical Engineering at the
University ‘Federico II’ of Naples (Italy). He completed his master thesis, entitled
‘Temperature effects in flow induced crystallization of thermoplastic polymers’ under the
supervision of Prof. Nino Grizzuti.
At the beginning of 2004, he started his Ph.D. in the Polymer Technology group
(SKT) of Prof. P.J. Lemstra at Eindhoven University of Technology under the guidance of
Prof. dr. S.Rastogi and dr. G.W.M. Peters.
135